Sunteți pe pagina 1din 756

ROCK QUALITY, SEISMIC VELOCITY, ATTENUATION AND ANISOTROPY

BALKEMA – Proceedings and Monographs


in Engineering, Water and Earth Sciences
ROCK QUALITY, SEISMIC
VELOCITY, ATTENUATION
AND ANISOTROPY

NICK BARTON

LONDON / LEIDEN / NEW YORK / PHILADELPHIA / SINGAPORE


Taylor & Francis is an imprint of the Taylor & Francis Group, an informa business

© 2007 Taylor & Francis Group, London, UK

This edition published in the Taylor & Francis e-Library, 2007.


“To purchase your own copy of this or any of Taylor & Francis or Routledge’s
collection of thousands of eBooks please go to www.eBookstore.tandf.co.uk.”

All rights reserved. No part of this publication or the information contained herein may be reproduced, stored in a retrieval system, or transmitted in
any form or by any means, electronic, mechanical, by photocopying, recording or otherwise, without written prior permission from the publishers.

Although all care is taken to ensure the integrity and quality of this publication and the information herein, no responsibility is assumed by the
publishers nor the author for any damage to property or persons as a result of operation or use of this publication and/or the information contained
herein.

Published by: Taylor & Francis/Balkema


P.O. Box 447, 2300 AK Leiden, The Netherlands
e-mail: Pub.NL@tandf.co.uk
www.balkema.nl, www.taylorandfrancis.co.uk, www.crcpress.com

Library of Congress Cataloguing-in-Publication Data

Barton, Nick, 1944–


Rock quality, seismic velocity, attenuation, and anisotropy/Nick Barton.
p. cm.
ISBN 0-415-39441-4 (hardcover: alk. paper)
1. Soil-structure interaction. 2. Earthquake engineering. I. Title.
TA711.5.B37 2006
624.151—dc22
2006005909
ISBN 0-203-96445-4 Master e-book ISBN

ISBN10: 0-415-39441-4 (Hbk)


ISBN13: 978-0-415-39441-3 (Hbk)
Table of contents

Preface XIII
Introduction XIX
The multi-disciplinary scope of seismic and rock quality XIX
Revealing hidden rock conditions XX
Some basic principles of P, S and Q XX
Q and Q XXI
Limitations of refraction seismic bring tomographic solutions XXII
Nomenclature XXIII

PART I

1 Shallow seismic refraction, some basic theory, and the importance of rock type 3
1.1 The challenge of the near-surface in civil engineering 3
1.2 Some basic aspects concerning elastic body waves 4
1.2.1 Some sources of reduced elastic moduli 5
1.3 Relationships between Vp and Vs and their meaning in field work 6
1.4 Some advantages of shear waves 7
1.5 Basic estimation of rock-type and rock mass condition, from shallow seismic P-wave velocity 9
1.6 Some preliminary conversions from velocity to rock quality 12
1.7 Some limitations of the refraction seismic velocity interpretations 13
1.8 Assumed limitations may hide the strengths of the method 16
1.9 Seismic quality Q and apparent similarities to Q-rock 17
2 Environmental effects on velocity 19
2.1 Density and Vp 19
2.2 Porosity and Vp 24
2.3 Uniaxial compressive strength and Vp 25
2.4 Weathering and moisture content 27
2.5 Combined effects of moisture and pressure 30
2.6 Combined effects of moisture and low temperature 32
3 Effects of anisotropy on Vp 35
3.1 An introduction to velocity anisotropy caused by micro-cracks and jointing 35
3.2 Velocity anisotropy caused by fabric 38
3.3 Velocity anisotropy caused by rock joints 40
3.4 Velocity anisotropy caused by interbedding 45
3.5 Velocity anisotropy caused by faults 47
4 Cross-hole velocity and cross-hole velocity tomography 49
4.1 Cross-hole seismic for extrapolation of properties 49
4.2 Cross-hole seismic tomography in tunnelling 52
4.3 Cross-hole tomography in mining 58
4.4 Using tomography to monitor blasting effects 61
4.5 Alternative tomograms 64
4.6 Cross-hole or cross-well reflection measurement and time-lapse tomography 66
VI Table of contents

5 Relationships between rock quality, depth and seismic velocity 69


5.1 Some preliminary relationships between RQD, F, and Vp 69
5.2 Relationship between rock quality Q and Vp for hard jointed, near-surface rock masses 74
5.3 Effects of depth or stress on acoustic joint closure, velocities and amplitudes 77
5.3.1 Compression wave amplitude sensitivities to jointing 83
5.3.2 Stress and velocity coupling at the Gjøvik cavern site 88
5.4 Observations of effective stress effects on velocities 88
5.5 Integration of velocity, rock mass quality, porosity, stress, strength, deformability 92

6 Deformation moduli and seismic velocities 97


6.1 Correlating Vp with the ‘static’ moduli from deformation tests 97
6.2 Dynamic moduli and their relationship to static moduli 104
6.3 Some examples of the three dynamic moduli 109
6.4 Use of shear wave amplitude, frequency and petite-sismique 110
6.5 Correlation of deformation moduli with RMR and Q 111

7 Excavation disturbed zones and their seismic properties 117


7.1 Some effects of the free-surface on velocities and attenuation 117
7.2 EDZ phenomena around tunnels based on seismic monitoring 119
7.3 EDZ investigations in selected nuclear waste isolation studies 124
7.3.1 BWIP – EDZ studies 124
7.3.2 URL – EDZ studies 127
7.3.3 Äspö – EDZ studies 131
7.3.4 Stripa – effects of heating in the EDZ of a rock mass 133
7.4 Acoustic detection of stress effects around boreholes 136

8 Seismic measurements for tunnelling 139


8.1 Examples of seismic applications in tunnels 139
8.2 Examples of the use of seismic data in TBM excavations 148
8.3 Implications of inverse correlation between TBM advance rate and Vp 149
8.4 Use of probe drilling and seismic or sonic logging ahead of TBM tunnels 151
8.5 In-tunnel seismic measurements for looking ahead of the face 152
8.6 The possible consequences of insufficient seismic investigation due to depth limitations 154

9 Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 159
9.1 Correlation between Vp and Lugeon value 159
9.2 Rock mass deformability and the Vp-L-Q correlation 162
9.3 Velocity and permeability measurements at in situ block tests 165
9.4 Detection of permeable zones using other geophysical methods 169
9.5 Monitoring the effects of grouting with seismic velocity 170
9.6 Interpreting grouting effects in relation to improved rock mass Q-parameters 172

PART II

10 Seismic quality Q and attenuation at many scales 181


10.1 Some basic aspects concerning attenuation and Q seismic 181
10.1.1 A preliminary discussion of the importance of strain levels 183
10.1.2 A preliminary look at the attenuating effect of cracks of larger scale 184
10.2 Attenuation and seismic Q from laboratory measurement 186
10.2.1 A more detailed discussion of friction as an attenuation mechanism 187
Table of contents VII

10.2.2 Effects of partial saturation on seismic Q 189


10.3 Effect of confining pressure on seismic Q 190
10.3.1 The four components of elastic attenuation 193
10.3.2 Effect on Q p and Q s of loading rock samples towards failure 195
10.4 The effects of single rock joints on seismic Q 197
10.5 Attenuation and seismic Q from near-surface measurements 202
10.5.1 Potential links to rock mass quality parameters in jointed rock 202
10.5.2 Effects of unconsolidated sediments on seismic Q 205
10.5.3 Influence of frequency variations on attenuation in jointed and bedded rock 207
10.6 Attenuation in the crust as interpreted from earthquake coda 209
10.6.1 Coda Q c from earthquake sources and its relation to rock quality Q c 209
10.6.2 Frequency dependence of coda Q c due to depth effects 210
10.6.3 Temporal changes of coda Q c prior to earthquakes 212
10.6.4 Possible separation of attenuation into scattering and intrinsic mechanisms 213
10.6.5 Changed coda Q during seismic events 214
10.6.6 Attenuation of damage due to acceleration 218
10.6.7 Do microcracks or tectonic structure cause attenuation 219
10.6.8 Down-the-well seismometers to minimise site effects 221
10.6.9 Rock mass quality parallels 224
10.7 Attenuation across continents 226
10.7.1 Plate tectonics, sub-duction zones and seismic Q 226
10.7.2 Young and old oceanic lithosphere 228
10.7.3 Lateral and depth variation of seismic Q and seismic velocity 228
10.7.4 Cross-continent Lg coda Q variations and their explanation 230
10.7.5 Effect of thick sediments on continental Lg coda 231
10.8 Some recent attenuation measurements in petroleum reservoir environments 232
10.8.1 Anomalous values of seismic Q in reservoirs due to major structures 235
10.8.2 Evidence for fracturing effects in reservoirs on seismic Q 236
10.8.3 Different methods of analysis give different seismic Q 238

11 Velocity structure of the earth’s crust 241


11.1 An introduction to crustal velocity structures 241
11.2 The continental velocity structures 244
11.3 The continental margin velocity structures 254
11.3.1 Explaining a velocity anomaly 256
11.4 The mid-Atlantic ridge velocity structures 261
11.4.1 A possible effective stress discrepancy in early testing 263
11.4.2 Smoother depth velocity models 265
11.4.3 Recognition of lower effective stress levels beneath the oceans 266
11.4.4 Direct observation of sub-ocean floor velocities 267
11.4.5 Sub-ocean floor attenuation measurements 268
11.4.6 A question of porosities, aspect ratios and sealing 270
11.4.7 A velocity-depth discussion 271
11.4.8 Fracture zones 272
11.5 The East Pacific Rise velocity structures 273
11.5.1 More porosity and fracture aspect ratio theories 276
11.5.2 First sub-Pacific ocean core with sonic logs and permeability tests 277
11.5.3 Attenuation and seismic Q due to fracturing and alteration 279
11.5.4 Seismic attenuation tomography across the East Pacific Rise 281
11.5.5 Continuous sub-ocean floor seismic profiles 283
VIII Table of contents

11.6 Age effects summary for Atlantic Ridge and Pacific Rise 287
11.6.1 Decline of hydrothermal circulation with age and sediment cover 289
11.6.2 The analogy of pre-grouting as a form of mineralization 291

12 Rock stress, pore pressure, borehole stability and sonic logging 295
12.1 Pore pressure, over-pressure, and minimum stress 295
12.1.1 Pore pressure and over-pressure and cross-discipline terms 295
12.1.2 Minimum stress and mud-weight 296
12.2 Stress anisotropy and its intolerance by weak rock 297
12.2.1 Reversal of Ko trends nearer the surface 299
12.3 Relevance to logging of borehole disturbed zone 301
12.4 Borehole in continuum becomes borehole in local discontinuum 302
12.5 The EDZ caused by joints, fractures and bedding-planes 306
12.6 Loss of porosity due to extreme depth 311
12.7 Dipole shear-wave logging of boreholes 312
12.7.1 Some further development of logging tools 315
12.8 Mud filtrate invasion 316
12.9 Challenges from ultra HPHT 320

13 Rock physics at laboratory scale 323


13.1 Compressional velocity and porosity 323
13.2 Density, Vs and Vp 324
13.3 Velocity, aspect ratio, pressure, brine and gas 326
13.4 Velocity, temperature and influence of fluid 328
13.5 Velocity, clay content and permeability 331
13.6 Stratigraphy based velocity to permeability estimation 332
13.6.1 Correlation to field processes 334
13.7 Velocity with patchy saturation effects in mixed units 335
13.8 Dynamic Poisson’s ratio, effective stress and pore fluid 337
13.9 Dynamic moduli for estimating static deformation moduli 339
13.10 Attenuation due to fluid type, frequency, clay, over-pressure, compliant minerals,
dual porosity 341
13.10.1 Comparison of velocity and attenuation in the presence of gas or brine 341
13.10.2 Attenuation when dry or gas or brine saturated 341
13.10.3 Effect of frequency on velocity and attenuation, dry or with brine 342
13.10.4 Attenuation for distinguishing gas condensate from oil and water 343
13.10.5 Attenuation in the presence of clay content 345
13.10.6 Attenuation due to compliant minerals and microcracks 346
13.10.7 Attenuation with dual porosity samples of limestones 348
13.10.8 Attenuation in the presence of over-pressure 350
13.11 Attenuation in the presence of anisotropy 351
13.11.1 Attenuation for fluid front monitoring 352
13.12 Anisotropic velocity and attenuation in shales 354
13.12.1 Attenuation anisotropy expressions ,  and  356
13.13 Permeability and velocity anisotropy due to fabric, joints and fractures 357
13.13.1 Seismic monitoring of fracture development and permeability 359
13.14 Rock mass quality, attenuation and modulus 365

14 P-waves for characterising fractured reservoirs 369


14.1 Some classic relationships between age, depth and velocity 369
Table of contents IX

14.2 Anisotropy and heterogeneity caused by inter-bedded strata and jointing 372
14.2.1 Some basic anisotropy theory 373
14.3 Shallow cross-well seismic tomography 374
14.3.1 Shallow cross-well seismic in fractured rock 377
14.3.2 Cross-well seismic tomography with permeability measurement 377
14.3.3 Cross-well seismic in deeper reservoir characterization 378
14.4 Detecting finely inter-layered sequences 379
14.4.1 Larger scale differentiation of facies 380
14.5 Detecting anisotropy caused by fractures with multi-azimuth VSP 382
14.5.1 Fracture azimuth and stress azimuth from P-wave surveys 382
14.5.2 Sonic log and VSP dispersion effects and erratic seismic Q 386
14.6 Dispersion as an alternative method of characterization 386
14.7 AVO and AVOA using P-waves for fracture detection 388
14.7.1 Model dependence of AVOA fracture orientation 391
14.7.2 Conjugate joint or fracture sets also cause anisotropy 392
14.7.3 Vp anisotropy caused by faulting 394
14.7.4 Poisson’s ratio anisotropy caused by fracturing 394
14.8 4C four-component acquisition of seismic including C-waves 394
14.9 4D seismic monitoring of reservoirs 397
14.9.1 Possible limitations of some rock physics data 397
14.9.2 Oil saturation mapping with 4D seismic 397
14.10 4D monitoring of compaction and porosity at Ekofisk 398
14.10.1 Seismic detection of subsidence in the overburden 400
14.10.2 The periodically neglected joint behaviour at Ekofisk 401
14.11 Water flood causes joint opening and potential shearing 402
14.12 Low frequencies for sub-basalt imaging 403
14.13 Recent reservoir anisotropy investigations involving P-waves and attenuation 404

15 Shear wave splitting in fractured reservoirs and resulting from earthquakes 407
15.1 Introduction 407
15.2 Shear wave splitting and its many implications 408
15.2.1 Some sources of shear-wave splitting 410
15.3 Crack density and EDA 411
15.3.1 A discussion of ‘criticality’ due to microcracks 412
15.3.2 Temporal changes in polarization in Cornwall HDR 413
15.3.3 A critique of Crampin’s microcrack model 415
15.3.4 90°-flips in polarization 415
15.4 Theory relating joint compliances with shear wave splitting 416
15.4.1 An unrealistic rock simulant suggests equality between ZN and ZT 417
15.4.2 Subsequent inequality of ZN and ZT 419
15.4.3 Off-vertical fracture dip or incidence angle, and normal compliance 419
15.4.4 Discussion of scale effects and stiffness 421
15.5 Dynamic and static stiffness tests on joints by Pyrak-Nolte 422
15.5.1 Discussion of stiffness data gaps and discipline bridging needs 424
15.5.2 Fracture stiffness and permeability 425
15.6 Normal and shear compliance theories for resolving fluid type 425
15.6.1 In situ compliances in a fault zone inferred from seismic Q 427
15.7 Shear wave splitting from earthquakes 428
15.7.1 Shear-wave splitting in the New Madrid seismic zone 428
15.7.2 Shear-wave splitting at Parkfield seismic monitoring array 429
X Table of contents

15.7.3 Shear-wave splitting recorded at depth in Cajon Pass borehole 432


15.7.4 Stress-monitoring site (SMS) anomalies from Iceland 432
15.7.5 SW-Iceland, Station BJA shear wave anomalies 433
15.7.6 Effects of shearing on stiffness and shear wave amplitude 435
15.7.7 Shear-wave splitting at a geothermal field 435
15.7.8 Shear wave splitting during after-shocks of the Chi-Chi earthquake in Taiwan 436
15.7.9 Shear-wave splitting under the Mid-Atlantic Ridge 436
15.8 Recent cases of shear wave splitting in petroleum reservoirs 438
15.8.1 Some examples of S-wave and PS-wave acquisition methods 438
15.8.2 Classification of fractured reservoirs 440
15.8.3 Crack density and shearing of conjugate sets at Ekofisk might enhance splitting 442
15.8.4 Links between shear wave anisotropy and permeability 445
15.8.5 Polarization-stress alignment from shallow shear-wave splitting 447
15.8.6 Shear-wave splitting in argillaceous rocks 450
15.8.7 Time-lapse application of shear-wave splitting over reservoirs 451
15.8.8 Temporal shear-wave splitting using AE from the Valhall cap-rock 454
15.8.9 Shear-wave splitting and fluid identification at the Natih field 455
15.9 Dual-porosity poro-elastic modelling of dispersion and fracture size effects 459
15.9.1 A brief survey of rock mechanics pseudo-static models of jointed rock 460
15.9.2 A very brief review of slip-interface, fracture network and poro-elastic crack models 461
15.9.3 Applications of Chapman model to Bluebell Altamont fractured gas reservoir 471
15.9.4 The SeisRox model 475
15.9.5 Numerical modelling of dynamic joint stiffness effects 476
15.9.6 A ‘sugar cube’ model representation 479
15.10 A porous and fractured physical model as a numerical model validation 480

16 Joint stiffness and compliance and the joint shearing mechanism 483
16.1 Some important non-linear joint and fracture behaviour modes 483
16.2 Aspects of fluid flow in deforming rock joints 486
16.2.1 Coupled stress-flow behaviour under normal closure 487
16.2.2 Coupled stress-flow behaviour under shear deformation 488
16.3 Some important details concerning rock joint stiffnesses Kn and Ks 492
16.3.1 Initial normal stiffness measured at low stress 494
16.3.2 Normal stiffness at elevated normal stress levels 495
16.4 Ratios of Kn over Ks under static and dynamic conditions 497
16.4.1 Frequency dependence of fracture normal stiffness 497
16.4.2 Ratios of static Kn to static Ks for different block sizes 498
16.4.3 Field measurements of compliance ZN 499
16.4.4 Investigation of normal and shear compliances on artificial surfaces in limestones 501
16.4.5 The Worthington-Lubbe-Hudson range of compliances 503
16.4.6 Pseudo-static stiffness data for clay filled discontinuities
and major shear zones 505
16.4.7 Shear stress application may apparently affect compliance 506
16.5 Effect of dry or saturated conditions on shear and normal stiffnesses 507
16.5.1 Joint roughness coefficient (JRC) 508
16.5.2 Joint wall compression strength (JCS) 509
16.5.3 Basic friction angle b and residual friction angle r 509
16.5.4 Empirical equations for the shear behaviour of rock joints 511
16.6 Mechanical over-closure, thermal-closure, and joint stiffness modification 513
16.6.1 Normal stiffness estimation 515
Table of contents XI

16.6.2 Thermal over-closure of joints and some implications 515


16.6.3 Mechanical over-closure 517
16.7 Consequences of shear stress on polarization and permeability 517
16.7.1 Stress distribution caused by shearing joints, and possible consequences for shear
wave splitting 518
16.7.2 The strength-deformation components of jointed rock masses 520
16.7.3 Permeability linked to joint shearing 523
16.7.4 Reservoir seismic case records with possible shearing 525
16.7.5 The apertures expected of highly stressed ‘open’ joints 526
16.7.6 Modelling apertures with the BB model 531
16.7.7 Open joints caused by anisotropic stress, low shear strength, dilation 534
16.8 Non-linear shear strength and the critical shearing crust 536
16.8.1 Non-linear strength envelopes and scale effects 536
16.9 Critically stressed open fractures that indicate conductivity 541
16.9.1 The JRC contribution at different scales and deformations 544
16.9.2 Does pre-peak or post-peak strength resist the assumed crustal shear stress? 545
16.10 Rotation of joint attributes and unequal conjugate jointing may explain azimuthal
deviation of S-wave polarization 548
16.11 Classic stress transformation equations ignore the non-coaxiality of stress and displacement 552
16.12 Estimating shallow crustal permeability from a modified rock quality Q-water 554
16.12.1 The problem of clay-sealed discontinuities 555

17 Conclusions 559

Appendix A – The Qrock parameter ratings 615


The six parameters defined 615
Combination in pairs 615
Definitions of characterization and classification as used in rock engineering 615
Notes on Q-method of rock mass classification 615

Appendix B – A worked example 625

References 627

Index 655

Colour Plates 721


Preface

This book traces an accelerating path through an important part of the earth sciences, describing seismic behaviour
and rock mechanics interpretation at many scales, to illuminate what lies beneath the earth’s immediate surface.
Although geophysics, and the rock mechanics and engineering geology of discontinuous media share the same
medium, they have had a mostly separate development – with little cross-referencing in the multitude of journals.
Regrettably, we seldom see geophysics colleagues at our rock conferences. This book attempts to bridge this void in
strategic locations.
Seismic velocity, seismic quality (the inverse of attenuation), and anisotropy are some of the very basics of geo-
physics, and they depend absolutely on the rock and fluid properties, the rock mass structures, the jointing, the frac-
turing, the microcracks and the other pore space. These are some of the fundamentals of earth science. All
contribute to the resultant dynamic stiffnesses, and to the fluid pressure micro-flow reactions, whether at dam foun-
dation depths, tunnel depths, reservoir-well depths, or earthquake depths. All components of the anisotropic,
dynamic, stiffness-velocity-permeability half-space, respond together in a logical pattern. Attempting to understand
this pattern is a major objective of this book.
The assumed ‘shared earth’ response is revealing itself with increasing speed. Despite the very small strains and dis-
placements involved in seismic wave loading there are inevitable, encouraging parallels, to the rock mechanics of larger
strains and displacements. This makes seismic response more understandable and more logical for a wider group of pro-
fessionals, with contributing areas of expertise.
In synthetic modelling in geophysics, there is now much interest in the rock joint or rock fracture compliances that
may hold part of the secret of fractured reservoir description. These same properties, when inverted, are used over much
larger displacements, in rock stability and deformation modelling. Remarkably, the dynamic compliance and static
stiffness of fractures and joints have mostly had a compartmentalized development in the different disciplines. A
dynamic, micro-strain-based normal compliance of 1013 m/Pa1 derived from shear-wave anisotropy measurement in
the sub-surface, is of recognisable magnitude when inverted, to compare with the pseudo-static ‘macro-strain’ joint nor-
mal stiffness (i.e. 10,000 MPa/mm or 10 MPa/micron) obtained from incremental loading tests on similar rock joints
at similar high stress levels.
The level of rock stress, the joint wall roughness, and the joint wall compressive strength, which are also important
components of aperture and permeability, provide estimates of these physical properties, not just the diagonal mem-
bers of a stiffness matrix. Here we have a classic reason for a disconnect between part of the earth sciences, which can be
bridged with advantage.
Attenuation and rock quality, another area of disconnect, can also be linked, but not quite so simply as taking the
inverse of attenuation and calling it seismic quality. The universally used seismic quality Q of geophysics, that we
will often call Q seis, shows some qualitative and quantitative connections to rock mass quality, also called Q, and
widely used in rock engineering since the 1970’s. The rock mass quality (Q), which we will often call Q rock, is com-
posed of several attenuation-causing parameters, that are directly equivalent to block size, inter-block friction and a
rough measure of effective stress and permeability.
There are clear, broad links between Q rock and Q seis, due to the discovery of a mutual connection to the empir-
ically derived and stress-dependent deformation modulus of rock masses. This connection is despite the fact
that only micro-strains, micro-displacements, and micro-flows (squirt) occur with the passage of dynamic waves.
Rock mass behaviour is non-linear and scale-dependent. Load-deformation curves have different gradients at differ-
ent stress levels. Dynamic waves seem to sense this non-linearity, and they apparently sense some of the scale effect
too.
This book is dedicated to making some of these cross-discipline empirical connections, in a simple non-mathe-
matical way, so that the people who see a lot of rock in their daily endeavours (geologists, engineering geologists,
rock mechanics and rock engineers), and those who see, and interpret, and model complex seismic results, from
XIV Preface

earthquakes, from fractured petroleum reservoirs, and from laboratory rock physics reservoir simulations, can more
easily communicate in the common anisotropic stiffness-velocity-permeability half-space that is earth science.
Communication in words and diagrams, rather than through complex formulae and matrices. At least half of the
people working in the earth sciences are not as good at mathematics as the other half may have assumed.

Acknowledgements

First and foremost this book is an acknowledgement to the many thousands of earth scientists working with geo-
physical interpretation of the near-surface, the sub-ocean, and the seismic shallow crust. Their dedication and inter-
esting publications have made this book a possibility. This volume is a well-illustrated documentation of just some
of their excellent work. The journey through their contributions has been one of increasing excitement.
Efforts have been made to reproduce the physical essence of reviewed work with suitable choice of author’s fig-
ures. Ricardo and Marcelo Abrahão have excelled in the expert redrawing of such figures, and are sincerely thanked
for their painstaking work. The writer’s summaries of key aspects of reviewed work are interspersed with personal
and rock mechanics based interpretations with which authors need not be in full agreement.
Material contributions, in the form of inaccessible articles, figures and data, and some valuable discussions and
improved insight, have kindly been provided by Dr. Enru Liu, Dr. Eda Quadros, Dr. Baotang Shen, Dr. Axel
Makurat, Prof. Stavros Bandis, Dr. Karstein Monsen, Prof. Michael King, Dr. Stuart Crampin, Dr. Heloise Lynn,
Harald Westerdahl, Dr. Sonja Maultzsch, Dr. Paul Chapman, Dr. Rudi Lubbe, Dr. Tor Arne Johansen, Dr. Barry
New, Dr. Saul Denekamp and Dr. Tore Lasse By, who enthusiastically introduced the writer to cross-hole seismic
tomography in 1986.
Part I of this book was mostly completed while the writer was Visiting Professor in the University of São Paulo
Polytechnic (USP). The writer’s kind neighbour in the Mining Department, Prof. Lineu Ayres da Silva, was indir-
ectly responsible for the five years extension involved in starting and completing Part II of this book. A recently pur-
chased volume by Kearey and Vine, 1996 lay open on his desk. A plate tectonics section of a plunging sub-ducting
crust with labels ‘low Q’, ‘high Q’ caught the writer’s rock-engineering attention. What did this ‘Q’ mean? Some of
the complex answers, and a simple one showing promise, will be found in Part II.
My final acknowledgements are firstly to Pat Coughlin, who has ensured a smooth-running and expert manuscript
production over a long period of endeavour. This started with the deciphering of handwriting and ended with countless
explanations of Microsoft’s hidden logic. The enthusiastic team at Taylor & Francis, Germaine Seijger and Lukas Goosen
and the Charon Tec team have produced a work to be proud of. The reader can be the judge of this. Finally my thanks
and apologies to a tolerant and loving wife Eda, who also ensured some key insights into rock-fluid interactions.

Permissions to Reproduce Figures

The nature of this book, specifically a wide-reaching literature review, involving some 830 references from some
forty different journals and publishing houses, has made obtaining permissions to reproduce figures a daunting and
sometimes impossible task regarding author-permissions, due to the several hundreds of first authors, and thousands
of multiple authorships. There are instances where we have been unable to trace or contact the copyright holder. If
notified, the publisher will be pleased to rectify any errors or omissions at the earliest opportunity. Many key authors
are retired, regrettably some have died, including Bengt Sjögren, who’s published work from 1979, 1984 and 2000
was an important source for key figures in several chapters of Part I. The most prominent authors have kindly given
permission for multiple reproduction of figures from my limited selection from their important contributions. All
publishers as listed below, have kindly given their permission for multiple reproduction of the numerous figures
reproduced in this reference volume. Their joint permissions, and those of contacted authors, and the contribution
of all authors that could not be contacted for whatever reason, are gratefully acknowledged. Their excellent work,
reproduced in this book, is a sincere acknowledgement of their contributions to geophysics.
Preface XV

Acoustical Society of America (ASA): Journal of the Acoustical Society of America: Figure 13.42
American Association of Petroleum Geologists (AAPG): Figure 15.36
American Geophysical Union (AGU): Journal of Geophysical Research: Figures 2.8, 3.1, 5.33, 5.34, 5.35, 10.14,
10.21, 10.25, 10.27, 10.28, 10.33, 10.37, 10.38, 10.41, 10.43, 10.44, 10.47, 10.48, 10.52, 10.53, 10.55,
10.58–10.60, 11.1, 11.6, 11.7, 11.8, 11.9ab, 11.10–11.21, 11.24–11.30, 11.31a, 11.32, 11.33, 11.35, 11.36,
11.38, 11.40–11.42, 11.48, 11.49, 11.52, 11.54–11.64, 11.66–11.71, 12.11, 12.22, 12.23, 13.2, 13.5a, 13.25,
13.29, 13.32, 13.33, 13.46, 14.16, 14.25, 14.26, 15.8, 15.11, 15.14, 15.18, 15.63. Figure Part II; Tables: 10.5,
10.6, 11.2, 15.2, 15.3, 16.5, 16.6
American Institute of Mining, Metallurgical and Petroleum Engineers (AJME): 16.42, 16.68
American Institute of Physics (AIP): Figure 10.21
American Physical Society (APS): Physical Review E: Figure 10.64
American Society of Civil Engineering (ASCE): Journal of Geotechnical Engineering: Figure 2.15
American Society of Mechanical Engineers (ASME): Transactions of the American Society of Mechanical Engineering:
12.6; Journal of Applied Mechanics: 2.9
Blackwell Publishing: Geophysical Prospecting: Figures 1.3, 1.5, 1.7, 1.8, 1.10, 1.11, 3.9, 4.3, 5.2–5.4, 5.10, 5.11,
6.11, 6.17, 8.12, 9.2, 10.65, 10.67, 13.24, 13.25, 13.36–13.41, 13.44, 13.48, 13.61, 14.15, 15.5, 15.6, 15.22, 15.28,
15.39, 15.40–15.42, 15.47, 15.48, 15.51–15.53, 15.55, 16.20–16.22; Geophysical Journal International (Geophys.
J. Int.): 10.22–10.24, 15.1a, 15.3, 15.4, 10.67; Other sources: Figures Part II, 11.1, 11.2, 11.18; Table 11.1
Cambridge University Press: Figures 11.3, 13.1, 13.2, 13.5 and 14.4
Centek Publishers, Luleå: Figure 16.13
Comité Francais de Géologie de l’Ingénieur et de l’Environnement (CFGI): Paris: Figures 5.6, 5.7, 8.5; Tables 8.1, 8.2
Coyne et Bellier: Figures 7.7, 6.19, 6.21
Elsevier: International Journal of Rock Mechanics and Mining Sciences and Geomechanics Abstracts: Figures 2.1,
3.2, 3.8, 4.7ab, 4.13, 4.14, 4.17, 4.20, 5.29, 5.30ab, 6.9, 6.20, 7.18, 7.20, 7.25, 7.26, 7.31, 7.32, 8.2–8.4, 9.6,
15.9, 13.53–13.55, 13.58, 15.17, 16.2, 16.4, 16.6, 16.7, 16.9, 16.12, 16.16, 16.17, 16.44, 16.46, 16.69, 16.73,
16.74; Table 4.1; Engineering Geology: Figures 5.17, 5.19, 15.26, 14.39; Journal of Applied Geophysics: 14.15,
15.5a, 15.56, 15.57; Table 16.8; Tectonophysics: Figures 11.31b, 11.53, 16.64, 16.65, 16.76. Other sources:
Figures 1.1, 1.6, 2.18, 4.12, 4.21, 5.13, 10.57, 11.5, 11.34, 15.23; Tables 2.2, 5.2, 11.1
European Association of Geoscientists and Engineers (AEGE): First Break: Figure 15.31; Other sources: Figures 10.2,
10.3, 10.10; 10.20, 10.21, 10.31, 10.36, 13.24, 14.37, 14.38, 15.27, 15.31, 15.37, 15.38, 15.43, 15.45, 15.46,
15.54; Table 13.2
Geophysical Research Letters: Figures: 4.9, 10.52, 11.51, 11.56, 15.44, 16.19; Other sources: Figures 9.7, 12.7,
12.8, 11.39, 11.46, 11.54, 13.11, Table 11.3
Geological Society of America (GSA): Geology: Figures 3.13, 10.6, 16.11, 16.56, 16.63; Figure 1.4
Geological Society: The Quarterly Journal of Engineering Geology: Figures 3.7, 3.10, 5.15, 5.16; Other sources: 2.12,
11.47, 13.56, 13.57, 15.16, 16.23
Imperial College, London: Figure 16.6
Imprime Adosa, Madrid: Figure 3.3, 5.1, 5.8, 5.9, 8.16
Institut du Bâtiment et des Travaux Publics; Annales d’ITBTP: Figure 6.20
Institut Français du Pétrole (JFP): Oil & Gas Science and Technology: Figures 3.5, 14.32, 14.33
XVI Preface

International Association for Engineering Geology and the Environment (IAEG): Figures 3.3, 5.1, 5.6–5.9, 8.5, 8.16;
Tables 8.1, 8.2, 16.7
International Commission on Large Dams (ICOLD), Paris: Various sources: Figures 3.6, 6.7, 7.4, 9.1, Table 6.3.
International Society for Rock Mechanics (ISRM): ISRM News Journal: Figures 7.2, 7.3, 6.18, 8.21–8.23, Table 12.1
Ishikawa Soil Incorporated Association: Figure 2.13
Japan Tunnelling Association (JTA); Tunnels and Underground: Figure 8.6
John Wiley & Sons: Figures 7.30, 6.23
Kansas Society of Petroleum Engineering: Figures 13.2, 13.5ab
Laboratório Nacional de Engenharia Civil (LNEC), Lissabon: Figures 2.2, 6.1, 6.15, 6.22, 6.23
Nagra; Nagra Bulletin: Figure 9.1
National Academy Press, Washington: Figures 6.2, 6.3, 6.8, 7.1, 7.8, 9.11; Table 6.2
Norwegian Petroleum Society (NFP): Figures 14.29, 14.30, 15.36
Office of Nuclear Waste Isolation (ONWI), Columbus: 16.10, 16.14, 16.15, 16.29–16.33, 16.46, 16.67
Österreichischen Gesellschaft für Geomechanik (ÖGG), Felsbau: Figure 6.4
Oyo Corporation: Figure 2.12
Royal Astronomical Society (RAS): Quarterly Journal of the Royal Astronomical Society: Figures 11.5, 11.37, 11.50
Schlumberger: Oilfield Reviews: Figures 12.24–12.26, 14.15, 15.1b, 15.19, 15.36, 15.1b; Other sources: 4.10
Seismological Society of America (SSA): Bulletin of the Seismological Society of America: Figures: 10.39, 10.40,
10.46, 10.52, 10.61, Table 10.7
SGE Editoriali, Padova: Figure 2.13
SKB, Stockholm: Figure 7.23
Society for Mining, Metallurgy and Exploration (SME): Various sources: Figures 2.4–2.7, 5.1a, 6.12, 7.12–7.15, 7.28,
15.7, 15.25, 16.27, 16.75
Society of Exploration Geophysicists (SEG): Geophysics: Figures 2.11, 2.19, 2.21, 3.11, 3.16, 4.15, 7.27, 10.1,
10.4–10.11, 10.13, 10.15ab, 10.16–10.19, 10.29, 10.30, 10.34, 10.35, 10.52, 10.64, 10.66, 10.68–10.72, 11.22,
11.23ab, 11.43, 11.48, 12.27, 13.3, 13.4, 13.6–13.8, 13.11–13.13, 13.17–13.23, 13.26–13.31, 13.34, 13.35,
13.42–13.45, 13.50–13.52, 14.1–14.3, 14.6–14.15, 14.18–14.24, 14.28, 14.31, 15.11, 15.29, 15.30, 15.60,
16.64. Tables: 10.10, 14.1–14.3, 15.1; The Leading Edge: Figures: 12.1a–d, 12.2a–d, 13.10, 13.14a–b, 13.15,
13.16, 14.33–14.36, 14.38, 15.15, 15.24, 15.35, 15.60; Canadian Journal Exploration Geophysics: Figures 10.63,
15.12–15.14, 15.32; Other sources: Figures 12.3, 12.30, 14.1, 15.5c, 15.10, 15.29, 15.44, 15.54, 15.65abc,
15.66, 16.6
Society of Petroleum Engineers (SPE): SPE Journal: Figures 13.2, 13.5ab, 14.32, 14.33; Other sources: Figures 12.12,
12.13, 12.29
Southern Africa Institute of Mining and Metallurgy (SIAMM): Figure 15.46
Springer Science and Business Media: Rock Mechanics: Figures 2.10, 16.10, 16.26, 16.41, 16.54ab, 16.57ab, 16.58;
Pure and Applied Geophysics – Pageophysik: 7.22ab, 10.12, 10.49ab, 10.50ab, 10.51, 10.52, 10.54; Other
Sources: Figures 10.42, 13.1, 16.60; Table 3.1
Stanford Rock Physics & Borehole Geophysics (SRB): Figures 13.2, 13.5ab
Preface XVII

Swedish National Science Council: Figure 1.45


Tapir Academic Press, Trondheim: Figure 15.25
Thomas Telford: Geotechnique: Figures 12.5, 12.9, 12.10, 15.2, 16.2, 16.8, 16.53, 16.75
Other sources: Figure 4.4, Tables 1.2, 1.3
University of California Berkeley: Figure 16.46
Wilmington: Tunnel & Tunnelling International: Figure 9.13

PhD Theses:
S. Bandis, 1980, University of Leeds (Fig. 16.3?, 16.16, 16.18, 16.40, 16.47, 16.52, 16.66, Tbl. 16.2, 16.3);
T. Cadoret, 1993, University of Paris (Fig. 13.2, 13.5ab, 13.20);
D. Han, 1986, Stanford University (Fig. 13.2, 13.4, 13.5e);
K. Iwai, 1976, University of California Berkeley (Fig. 16.46);
D.L. Jizba, 1991, Stanford University (Fig. 13.5d);
Y.-Q. Liu, 2003, University of Edinburgh (Fig. 14.15b);
R. Lubbe, 2005, Oxford University (Fig. 16.20, 16.23);
N. Lucet, 1989, University of Paris (Fig. 13.2, 13.5ab);
E. Quadros, 1982 (Msc), University of São Pualo (Fig. 16.6);
A. Shakeel, 1995, Imperial College, Univ. London (Fig. 13.58);
J.C. Sharp, 1970, University of London (Fig. 16.6);
C. Slater, 1997, University of Edinburgh (Fig. 15.20, 15.34, 15.35);
S.R. Tod, 2002, University of Cambridge (Fig. 15.44);
J. Yan, 2003, University of Edinburgh (Fig. 13.14);
J. Yuan, 2001, University of Edinburgh (Fig. 14.27).
Introduction

The multi-disciplinary scope of seismic and rock quality

Seismic, sonic and ultrasonic measurements are utilised by a large number of geo-science, geo-engineering and geo-
resource disciplines. Their use is so widespread in the earth-sciences, that it should be of no surprise to us that such
techniques are also used to register such diverse subjects as osteoporosis in cows, and the control of ‘crispiness in
breaded fried chicken nuggets’. The latter was a thesis in Biological Systems Engineering.
Since rock engineers tackle different problems from petroleum engineers and geophysicists, who in turn tackle
different problems from tectonophysicists, there has been an understandable yet regrettable compartmentalisation
between the disciplines. Both practitioners and researchers in each of these major fields, generally go to different
conferences and read and publish in different journals, as there are ‘too many’ choices of each in each discipline, even
in each speciality where we earn our living. The luxury of cross-discipline interaction, occasionally experienced with
great interest and resulting stimulation, is usually defeated by time, cost and also in part, by technical-language bar-
riers, and even mathematics.
An interesting example of partial ‘compartmentalization’ is stiffness and compliance. Each have followed almost
separate development since the late 1960s in rock mechanics, and since the early 1980s in geophysics. Each are
essential to each subject; for numerical modelling of stability and deformability in rock engineering; for improved
interpretation of attenuation, anisotropy and shear wave splitting in the geophysics of fractured petroleum reser-
voirs. Yet the dynamically measured, micro-deformation fracture compliances in geophysics (in the normal and
shear directions), are numerically close to the inverse of incrementally-loaded joint stiffnesses in rock mechanics, at
least when rock quality is high.
The frequently illustrated material in this book has been assembled as a result of an interest in a variety of civil,
mining, petroleum, geophysics and earth-science fields. The common denominator has been rock mass and rock
joint behaviour as presumably impacting the seismic interpretation. An interesting and very large selection of seis-
mic velocity and seismic quality related data, from practitioners working in widely varied disciplines, has been
assembled. Much has obviously been left out or not yet seen. Much is still under development.
The chapters of Part I are mostly civil engineering related with strong links to the interpretation of rock condi-
tions at both laboratory and field-scale, with their impact on engineering of tunnels and dams and planned nuclear
waste repositories. The chapters of Part II go deeper both figuratively and literally, and consider much larger scale
uses of seismic attributes, from hydrocarbon reservoirs and the use of multiple dynamic energy sources, to the inter-
pretation of mid-ocean spreading-ridges, to crustal conditions interpreted from natural earthquake hypocentres.
The phenomen of seismic anisotropy, known already in the nineteenth century to give lower stiffness perpen-
dicular to layering than parallel, is now in widespread use for investigating fractured rock at depth. Features of the
rock mass, though of sub-seismic-wave size, can be detected at many kilometers depth, due to shear wave splitting,
giving polarization parallel and perpendicular to dominant jointing. Different time delays for the fast and slow shear
wave components vary with fracture properties and with frequency, giving frequency-dependent anisotropy.
Efforts have been made to seek out and to reproduce in brief, with helpful figures, the seismic measurements and
interpretations which have a clear or potential rock quality content, at whatever scale. Clearly the term ‘rock quality’
conceals various techniques and scales of measurement, and varied interests in ‘rock quality’ per se. A rock mass with
high velocity and high rock quality (i.e., exhibiting low attenuation) would make life less profitable for machine bored
tunnellers due to slow progress and frequent cutter-changes. Aggregate producers would need more drilling and
explosives per ton, and would seek other quarries. The very existence of hydrocarbon reservoirs and their product-
ivity would be severely prejudiced if either ‘rock quality’ or ‘seismic quality’ was too high. Others would welcome
good ‘rock quality’ characteristics, for example producers of dimension stone and clients expecting cheap drill-and-
blasted tunnels requiring little rock support.
XX Introduction

Revealing hidden rock conditions

At the beginning of most rock engineering projects we are operating ‘blind’, and any help to ‘see’ what may lie below
our dam foundation, or ahead of our tunnel, saves schedules, budgets and sometimes lives as well. The beauty of
seismic, sonic or ultra-sonic investigations is that they can be applied over a virtually unlimited range of scales, to
‘see’ micro-cracks closing under stress in the laboratory, or to ‘see’ fluctuations in effective stress across a regional fault
caused by changes of reservoir level, and to monitor the effects of water-flooding in a fractured petroleum reservoir.
Already in 1917, Fessenden had proposed (and patented) the use of a cross-hole seismic technique to locate ore
bodies. The scale of investigation can be increased by orders of magnitude to ‘earth-scale’, when illuminating the
seismic structure of the earth’s crust, and further again to depths of 5000 km or more, to the solid iron core of the
earth, as a result of global-station analyses following large earthquakes.
Sjøgren, 1984, gave the civil engineering (near-surface) profession a particularly useful guide in the use of shallow seis-
mic refraction techniques for those involved in shallow sub-surface projects. The fundamental principle is that seismic
waves propagate with significantly different velocities in different near-surface geotechnical and geological strata, due to
the seismic visibility of weathered, low-stressed materials in general. This also means that the velocities tend to increase
rapidly with depth, which must not be misinterpreted as meaning better quality per se. Intermediate high-speed layers,
or hidden low velocity layers obviously disturb this simplified picture, and velocity anomalies and incorrect depth inter-
pretations result unless separate analysis i.e., downhole vertical seismic profiling (VSP), or coring is performed.
Fundamental difficulties in the context of rock engineering (and in all other disciplines too) are that the means of
access, superficial or along boreholes, are often limited by the geometry of the problem, by the (urban or sub-sea)
location, and by the cost. The freedom to choose optimal experimental layouts is therefore limited. As pointed out
by Cosma, 1995, this may cause blind zones, even in the immediate vicinity of the observation points.
In the case of soil or weathered rock horizons, seismic velocity interpretation readily distinguishes the water table
from a lithological boundary by inspection of the shear or transverse wave velocity (Vs). If this remains constant
across the region of changing water content, while Vp changes, a groundwater surface is indicated, since the shear
waves do not respond to changing water content due to the lack of shear stiffness. If Vs also changes, a geotechnical
or geological layer will have been crossed. Typical ranges of Vp for a variety of near-surface sediments and rocks are
reviewed in Chapter 1.
One of the historic and important applications of refraction seismic in civil engineering, has been at dam sites,
which were investigated in great numbers, especially in the 1950s, 60s and 70s. Rock quality, permeability, and
deformation modulus were of fundamental importance. Associated hydropower tunnels such as headrace and tail-
race tunnels have been the subject of countless thousands of kilometres of seismic refraction spreads, not to men-
tion all the power house foundations and high pressure penstock locations.
The seismic spreads at the ground surface should if possible be set out in optimal directions to investigate sus-
pected sub-surface anomalies. Since the ray paths are essentially following sub-horizontal paths, steeply dipping or
vertical features such as faults or deeply weathered zones can be readily located and given a characteristic seismic sig-
nature. Localised P-wave velocities of 2 or 3 or 5 km/s have distinct engineering implications for near-surface tun-
nelling or foundation stripping. Their interpretation in relation to rock type (uniaxial strength and porosity) and in
relation to the depth of measurement, or to stress level and stress-induced anisotropy, will be reviewed in detail in
this book, with the help of a quantitative rock mass quality description.

Some basic principles of P, S and Q

The P-wave is a longitudinal wave, in which the direction of particle motion coincides with the wave propagation.
It is often termed the first arrival or compressional wave. By contrast, the lower velocity transverse S-wave has par-
ticle motion in the plane perpendicular to the direction of wave propagation. An S-wave is of two possible basic
types: the SH-wave in which particle motion is parallel to a boundary, usually the ground surface, and the SV-wave
which has particle motion perpendicular to both the wave propagation direction, and to the particle motion of the
SH-wave.
Introduction XXI

When passing through anisotropically fractured petroleum reservoirs, a shear wave will likely split into fast (qS1) and
slower (qS2) polarized components, giving clues about the fracturing character and perhaps the principal stress direc-
tion. The latter coupling may be more complex than convention suggests however, due to adverse stress-closure-per-
meability behaviour in reservoir rocks, unless they are strong enough to tolerate tens of megapascals of effective
normal stress across their ‘open’ joints or fractures. Slight shearing and dilation may actually be needed on conju-
gate joint or fracture sets, to explain permeability and production from fractures in weaker reservoir rocks, and to
explain the ‘surprising’ maintenance of permeability deep into the crust.
There is a ‘problem’ of frequency dependence for all the component velocities of P- and S-waves, but in fact in the
problem lies the more accurate interpretation. There are exciting current developments in these dispersive, frequency-
dependent interpretations of velocities and attenuation, and in their relation to anisotropy, where rock mechanics
knowledge of ‘joint stiffnesses’, or their dynamic micro-strain-based near-inverses: the geophysicist’s ‘fracture com-
pliances’, are proving extremely important supplements to the earlier focus on the elliptic aspect ratios of micro-
cracks, and the larger-scale – and smaller magnitudes – of the aspect ratios of almost closed fractures.

Q and Q

Seismologists have had a long tradition of utilising a quality factor Q-seismic (with numerous sub-sets such as the
basic Qp, Qs, and Qc , the latter from the coda or tail-end of a dynamic wave sequence). Q-seismic was popularized
by a famous Knopoff, 1964 paper with the briefest possible title: ‘Q’. We will see the possibility of a Q-seismic rela-
tion with another quality descriptor called the ‘Q-value’, from rock engineering, not directly, but via a mutual
apparent relation to the stress-dependent pseudo-static deformation modulus: surprisingly not to the dynamic modu-
lus, at least not in the top kilometre or so.
Q-seismic is a dimensionless factor whose inverse (Q 1 seis) indicates, if simply stated, the percentage loss of energy
of a single wave length due to various (and sometimes disputed) mechanisms of attenuation in the rock mass at
many possible scales. Reduction in wave amplitude is the most obvious effect. The attenuation is caused by scatter-
ing from geo-structures of different scales, and by absorption in intrinsic micro-mechanisms like normal and shear
micro-displacements across microcracks and joints, therefore involving friction to some degree, and relative micro-
movement of fluids between the pore-space, the micro-cracks and the jointing or fracturing.
As a result of the passage of the very slightly deforming seismic waves there will be a lot of references to ‘squirt
flow’ losses in Part II of this book, in connection with anisotropic attenuation, which is one of several properties of
the fluid conducting structures of fractured or naturally jointed hydrocarbon reservoirs.
In parallel but previously almost unrelated endeavours, a prominent engineering geologist (Deere, 1964) developed
a simple empirical rock quality factor RQD, related with the degree of jointing or fracturing in drill-core. In the
1970s, with no knowledge of Qseis, the rock quality Q-value was developed, which includes RQD as one of the six
parameters. The rock engineering rock quality Q-value describes the degree of jointing (as relative block size) and
important ‘internal’ joint properties like roughness and clay-filling (giving the inter-block friction coefficient). It
also incorporates estimates of the permeability and the stress-to-strength ratio.
Frequent use will be made of the Barton et al., 1974 and Barton 2002 rock quality Q-value and Qc-value in vari-
ous parts of this book. It provides a simple link to seismic velocity, and it probably has the potential for explaining
some attenuation mechanisms as well. The rock quality Q-value has a six orders of magnitude scale of quality (from
0.001 to 1000), and it predicts a two to three orders of magnitude range of deformation modulus. Completely
unjointed, massive rock masses, with Q  1000, will clearly show almost no attenuation. At many kilometres
depth, Qseis values are of similar magnitude. Completely decomposed, near-surface, faulted rock with Q 
0.01–0.001 will obviously give complete attenuation (i.e. effectively lower than the theoretically lowest possible
Qseis and highest possible Q 1 seis – each probably beyond measurement limits).
It is expected that future graphs of Q (seismic quality factor) versus Q (rock quality factor) in rock masses (as
opposed to lab-samples), can show strong correlations in the future, when geophysics data is reported in parallel
with rock quality data. Each of the ‘Q-factors’ will be described in greater detail later in this book. We will also see
the ‘problem’ of frequency-dependence, and the ‘problem’ of anisotropy, but both these problem areas are obviously
XXII Introduction

concealing the potential for improved interpretation of the structures beneath the earth’s surface, both shallow and
deep.

Limitations of refraction seismic bring tomographic solutions

Refraction seismic methods have been used for at least fifty years, but have some fundamental limitations that include
masking of lower velocity layers under higher velocity materials, such as basalts above petroleum reservoirs, and dif-
ficulties with multiple velocity layers in close proximity. Some unexpectedly costly tunnelling has resulted from mis-
takes in interpretation, due to such features. However, as with most limitations, there are various solutions, and
geophysicists have been extremely creative, and also willing to modify and apply techniques from other well-funded
fields like medicine.
While P-wave and S-wave measurement between two points can be expressed as average wave velocities (or give
a rather unhelpful ‘average’ picture of a patients brain), there is the possibility of using more comprehensive mul-
tiple source and receiver positions in separate multiple-boreholes, thereby giving positional (2D or 3D) tomo-
graphic imaging. A tumour in an unfortunate patient, and real-time scanning of brain-wave activity, as illuminated
in medicine, have their engineering-scale equivalents. A fault zone delaying a tunnel, and four-dimensional fluid-
migration-imaging in a producing reservoir would be approximate, large-scale geophysics equivalents. The most
basic imaging analogy has been practiced for many years by geophysicists, who use earthquake sources and global
monitoring stations to deduce the structure of the whole earth. So perhaps geo-physicists actually helped to inspire
medical imaging of the human body?
In intermediate-scale, near-surface civil engineering, the strategic positioning of pairs of boreholes across complex
zones or faults can be used for optimal characterization of these features, if they appear to be a threat to progress of a
tunnel, or to dam foundation integrity. In special cases cross-hole tomography measurements may lead to the avoid-
ance of collapse, as more reliable decisions can be taken concerning the need for strengthening by pre-grouting, or the
need for special pre-installed ground support, or perhaps even ground-freezing. Tunnels with inadequate overburden
or severe water leakage potential such as inundation by rivers or lakes, or local inflows that would allow pore pressure
draw-down compaction in soft clays beneath important buildings, can also benefit greatly from seismic-based deci-
sions for special treatment of the ground.

Part I which occupies the first third of this book, will be found to contain mostly civil-engineering and tunnel engi-
neering treatments of the velocity-quality links that are helpful when interpreting near-surface conditions.The com-
plementary laboratory testing that has often accompanied geophysics investigations of the near-surface, will also
have emphasis on lower stress. Because of this, the effect of weathering and alteration and excavation on seismic
attributes, will each be emphasised. Despite the obvious challenges of seismic interpretation in fractured and faulted
petroleum reservoirs at many kilometers depth, or of mid-ocean ridge investigations beneath three kilometers of
ocean, many geophysicists insist that obtaining high resolution images from ground level to just 50 m depth, is still
one of the major challenges of modern geophysics. This happens to be the layer of the subsurface closest to most of
our civil engineering endeavours, from tunnels, to dams, to the foundations for high buildings.

Part II of this book tackles greater depths, greater scales, and more subtle geophysical detail, as benefits this rapidly
developing field. Geophysics has been in ‘rapidly developing’ phases many times in the past. The latest phase is due
to many parallel developments, not least an acceptance of the benefits of three-dimensional surveys, of monitoring
reservoir changes over time (4D), each requiring the ever-developing power of modern computers for the complex
processing of huge amounts of digital data. Investment in geophysics is growing further, due to the inestimable advan-
tages of improved information. The continued search for reliable earthquake precursors, and the pressures to find
more hydrocarbons in more heterogeneous reservoirs, and improve the recovery from those already being depleted,
are each driving the developments in this remarkable field. In the future, more geophysical investments may also be
used to aid in the search for potable water, which already far exceeds the price of gasoline in many locations.
Nomenclature

 angle subtended between a discontinuity and the major principal stress s1


 rock mass density (t/m3)
 shear-wave anisotropy parameter
 change in value (e.g. e, E applying to changes in joint or fracture apertures)
v vertical component of deformation
h horizontal component of deformation
m1 frequency of joints (or fractures) per meter (also F m1)

shear modulus
c uniaxial compression strength (MPa)
h min minimum horizontal component of stress
H max maximum horizontal component of stress
r radial stress around an excavation in rock
v vertical component of principal stress
1 2 3 principal stresses
tangential stress around a (circular) opening
max maximum tangential stress
min minimum tangential stress
shear stress (in a direct shear test)
 friction angle of joint, fracture, filled discontinuity, fault (geomechanics)
 fractional porosity (rock physics)
b basic friction angle, flat unweathered surfaces, low stress
c critical state line defining s1 = 3s3
peak peak friction angle of a joint, fracture
r residual friction angle of a joint, fracture, fault
 axial modulus
ANDRA Agence Nationale pour la gestion des Déchets Radioactifs
AR advance rate (TBM, actual weekly, monthly rate)
AVO amplitude variation with offset
AVOA amplitude variation with offset and azimuth
BB Barton-Bandis constitutive model for rock joints, used with UDEC as UDEC-BB
BEM boundary element method of numerical modelling
BGS British Geological Survey
BHA bottom hole assembly
BHC borehole compensated sonic logging tool
BHTV borehole televiwer
BISQ Biot and squirt flow model
BP British Petroleum
BWIP Basalt Waste Isolation Project, Hanford, Washington, USA
c cohesion of intact rock, joint, fracture, or rock mass
CBTF Conoco Borehole Test Facility
CC cohesive component of rock mass (from Q-value)
CDR compensated dual resistivity log
CSFT coupled stress/shear flow test/temperature, for HM, HTM testing of joints
XXIV Nomenclature

CSM Colorado School of Mines, Idaho Springs experimental mine facility


md,
d Darcy-based unit of permeability (md,
d for milli-, micro-darcies)
DEM distinct element modelling
D-H-M dynamic-hydraulic-mechanical coupling
e change of hydraulic aperture (joint, fracture: interpret from flow test)
E change of (mean) physical aperture (joint, fracture: interpret from deformation)
Edyn dynamic axial or Young’s modulus from VP and VS measurement
Ee modulus of elasticity (pseudo-static unloading stiffness: plate load test)
EF dyn (as Edyn but field-scale, based on seismic measurements, shortened to EF)
FL dyn (as Edyn, lab-scale, based on ultrasonic measurements, shortened to EL)
Emass pseudo-static modulus of deformation (also D, Ed and M) from loading stiffness of rock mass
e hydraulic aperture of a joint or fracture (kintrinsic laminar flow, defined as e2/12)
E mean physical aperture of joint or fracture (empirical JRC-estimated, or BB-model)
EDA extensive dilatancy anisotropy
EDZ excavation disturbed/damaged zone ( typically around tunnels)
Mini-EDZ ‘alteration zone’ typically around boreholes or wells
EOR enhanced oil recovery
F m1 frequency of fractures (or joints) per meter
FEM finite element method of numerical modelling
FC frictional component of rock mass (from Q-value)
FLAC two-dimensional continuum code for modelling small or large deformations in rock or soil
FLAC3D 3D continuum code for modelling small or large deformations in rock or soil
FM, FMS formation micro-scanner
FRACOD fracture mechanics boundary element code for modelling fracturing process in rock
FZI flow zone indicator
GRM generalized reciprocal method
HDR hot dry rock
HPHT high pressure high temperature (well)
HRSN high resolution seismic network, Parkfield, California
HSP horizontal (in-tunnel) seismic reflection profiling
HTI as TIH, transversely isotropic, horizontal axis of symmetry
HTM hydro-thermal-mechanical (coupling) (also MHT)
i with  or  implies dilation or contraction when loaded in shear
I50 point load index for 50 mm size samples
IPT Institute of Technological Research (S~ao Paulo)
ISONIC sonic while drilling tool
ISRM International Society of Rock Mechanics
Ja rating for joint alteration, discontinuity filling in Q-calculation
JCS joint wall compression strength (MPa)
Jn rating for number of joint sets in Q-calculation
Jr rating for joint surface roughness in Q-calculation
JRC joint roughness coefficient (dimensionless: range 0 to 20)
Jv volumetric joint count (sum of frequencies for different sets)
Jw rating for water softening, inflow and pressure effects in Q-calculation
K,k permeability (intrinsic: units of length2, engineering: units of m/s)
K bulk modulus (also Kbulk)
Kint intermediate principal permeability
Kmax maximum principal permeability
Kmin minimum principal permeability
Kn normal stiffness (of joint or fracture: strongly non-linear, sample dependent)
Nomenclature XXV

Kn dyn dynamic normal stiffness (of joint or fracture)


Ko ratio of rock stresses sh min/sv
KS shear stiffness (of joint or fracture: non-linear, sample dependent, scale dependent)
Ks dyn dynamic shear stiffness (of joint or fracture)
L Lugeon unit of water injection (l/min/m of borehole/1MPa excess pressure  107 m/s)
Lg coda waves, tail of seismogram
LOFS life of field seismic
LSS long-spaced sonic tool
LWD logging while drilling
M deformation modulus (pseudo-static loading stiffness: plate load test. Also Emass, D)
M1,2 dynamic elastic moduli at frequencies f1 and f2
MAR mid-Atlantic ridge
MHF massive hydraulic fracturing
MIT Massachusets Institute of Technology
MPBX multiple position borehole extensometer
MWD measurement while drilling
n effective stress coefficient (Biot)
n% porosity of matrix
NAFZ North Anatolian Fault Zone, Turkey
ND natural directivity
NGI Norwegian Geotechnical Institute, Oslo, Norway
NMO normal moveout
NPF Norsk Petroleumsforening (Norwegian Petroleum Society)
OC over-closure of joints, mechanical or thermal
O/R open/rock-to-rock sections of shearing joint, opposite rotation
OBC ocean bottom cable
OBS ocean bottom seismometers
P volumetric stress
Pg direct (P-) wave (crustal scale studies)
Pn refracted (P-) wave (crustal scale studies)
Pr support pressure, radial capacity of support in a tunnel
PR penetration rate (TBM, uninterrupted boring)
Q rock mass quality rating (‘Q-value’ range 103 to 103, dimensionless)
Qrock rock mass quality rating, distinguish from Qseis, seismic quality, inverse of attenuation
Qc seismic quality of coda wave
QE seismic quality in extensional resonance mode
Qe seismic quality component (Young’s mode of elastic excitation)
Qk seismic quality component (bulk mode of elastic excitation)
Qo seismic quality, Lg coda at 1 Hz
QP seismic quality of P-wave (through given medium)
Qs seismic quality component (shear mode of elastic excitation)
QS seismic quality of S-wave (through given medium)
Qc rock mass quality rating (Q or Qrock normalized by c/100)
Qo Q (or Qc or Qrock) calculated with RQDo, oriented in the loading or measurement direction
Qseis, seismic quality factor (‘Q’), inverse of attenuation, also for QP or QS, or the coda wave Qc
Qtbm rock-machine quality factor for TBM tunnel boring machines based partly on Q-value
QVO Q(seismic) versus offset
r,R Schmidt hammer rebound % on wet joint surfaces, dry intact samples, respectively
REV representative elemental volume
RMR rock mass rating developed by Bieniawski
XXVI Nomenclature

RQD rock quality designation developed by Deere (modified core recovery %)


RQDo RQD oriented in the loading or measurement direction
RQI reservoir quality index
SAFZ San Andreas Fault Zone
SCV Site Characterization and Validation, SKB project in Stripa mine, Sweden
S(fr) steel fibre reinforced sprayed concrete
S/C ratio ratio of subsidence to compaction magnitudes, above and within reservoirs
SKB Swedish Nuclear Fuel Co. (Stockholm)
SRF rating for faulting, strength/stress ratios, squeezing, swelling: 6th parameter in Q-value
TBM tunnel boring machine
3DEC three-dimensional distinct element code for modelling jointed rock masses
TIH transversely isotropic, horizontal axis of symmetry (also HTI)
TIV transversely isotropic, vertical axis of symmetry
TSP (in-tunnel) seismic reflection profiling
TSX tunnel sealing experiment
UCS uniaxial compressive strength of rock cylinder
UDEC universal distinct element code, for modelling jointed, fractured rock in 2D
(3DEC) three-dimensional distinct element code, for modelling jointed, fractured rock in 3D)
URL Underground Research Laboratory, Manitoba, Canada
Vp P-wave seismic velocity (km/s)
VS S-wave seismic velocity (km/s)
VSP vertical seismic profiling
WAP wide aperture profile
WIPP Waste Isolation Pilot Plant, New Mexico
w.r.t. with respect to (index only)
ZEDEX Zone of EXcavation Disturbance Experiment, SKB project, Äspö, Sweden
ZN dynamic compliance (of joint or fracture) ( 1/Kn dyn)
ZT dynamic compliance (of joint or fracture) ( 1/Ks dyn)

Cross-discipline differences and connections

• effective stress  total stress minus pore pressure in geomechanics


• differential stress  shear stress caused by 1 – 3 application in geomechanics
• differential pressure  confining pressure minus pore pressure in rock physics
• compliance  (dynamic stiffness)1, compliance  (pseudo-static stiffness)1
• Qseis  1/attenuation, Qrock  Qseis, but Qrock provides estimate of Emass  Qseis
Part I
1 Shallow seismic refraction,
some basic theory, and the
importance of rock type

‘Nature has left us an incomplete and often well-concealed record of her activities, and no ‘as con-
structed’ drawings!’ (Stapledon and Rissler, 1983)

‘Tenders for the Tay pipeline crossing did not allow time for boreholes to locate bedrock. Seismic
refraction took one day to confirm that the trench would not encounter rock. The pipeline was laid
in sediments.’ (Gardener, 1992)

‘The time may come when the various relations between geophysical parameters and rock properties
can be usefully combined into a single classification system.’ (Darracott and Orr, 1976)

1.1 The challenge of the that are perpendicular to the major stress, and the more
near-surface in civil open state of those that are parallel will give the rock
engineering mass anisotropic stiffness. Consequently the rock mass
will frequently display anisotropic seismic velocities. By
Refraction seismics is by far the oldest method used in implications, hydraulic conductivities and deformation
exploration seismology, with its origin traced to R. Mallet moduli that show anisotropic distributions will be, at least
from 1848. Shallow refraction seismic measurements in part, detectable by seismic measurements. Anisotropy
using first arrival, compressional P-wave velocities close will also be caused by layered inter-beds, foliation and
to the surface often give a remarkable picture of near schistocity, and of course by a dominant joint set. Simple
surface conditions due to some fortuitous interactions of examples of (azimuthal) anisotropy, applicable in civil
physical phenomena. Firstly, weathering and the usual engineering, will be given in Chapter 3, while larger-
lack of significant stress near the surface has allowed joint scale examples of anisotropy detection will be described
systems, shear zones and faults to be exaggerated in both in much greater detail, and from various fields of the
their extent and severity. Secondly, stress levels are low earthsciences, in Chapters 13, 14 and 15 in Part II.
enough to allow joints and discontinuities to be seismic- Despite the obvious challenges of seismic interpret-
ally visible due to their measurable apertures. ation in fractured and faulted petroleum reservoirs at
So-called acoustic closure occurs at greater depths than many kilometers depth, or of mid-ocean ridge investi-
those usually penetrated by conventional hammer seis- gations beneath three kilometers of ocean, many geo-
mic, unless rock strengths are rather low (e.g., New and physicists insist that obtaining high resolution images
West, 1980; Hudson et al., 1980). (At this juncture, we from ground level to just 50 m depth, is still one of the
need to differentiate between two ‘J.A. Hudson’ authors, major challenges of modern geophysics. This happens to
one in geophysics, the other in rock engineering, and be the layer of the subsurface closest to most of our civil
both very prominent in their chosen fields. We will engineering endeavours, from tunnels, to dams, to the
occasionally refer to ‘rock’ Hudson in Part I, and later foundations for high buildings.
in Part II to ‘seismic’ Hudson). Undoubtedly, the ‘0 to 50 m’ challenge is mainly due
Since micro-fractures and rock joints are sensitive to to the extreme variability of the near-surface, resulting
stress levels, the more closed state of the discontinuities from the contrasting geological materials and weathering
4 Rock quality, seismic velocity, attenuation and anisotropy

grades that are often present. There is also a velocity Sjøgren, 2000 suggested the following list of essen-
gradient that is extreme compared to anything found at tial information expected from near-surface seismic
greater crustal depths, where consolidation effects smooth surveys, performed for civil engineering geotechnical
out some of the differences. The first 5 m of unconsoli- investigations:
dated dry beach sand may see velocity increase from
150 m/s to 300 m/s, (Bachrach et al., 2000), giving a gra- ● The velocities of the overburden layers, including
dient of 30 s-1, which may be an order of magnitude the upper, less consolidated rock layers.
higher than the gradient over the next 50 to 100 m, where ● The thickness of the various overburden velocity
weathered and jointed rock may typically be found. layers, and the total depth to the main refractor.
There are an infinite number of challenges in the near- ● A detailed determination of the velocity distribu-
surface. Some of the worse may be karst phenomena in tion in the main refractor.
limestones, or the ‘inverse’ problems of core-stone anom- ● An estimate of the uncertainty of the velocity and
alies in the case of sparsely jointed but deeply weathered depth determinations.
granites and gneisses. These features have caused tun- ● An analysis of the (velocity-) depth structure.
nelling surprises in numerous countries, with nearly as ● An assessment of velocities in vertical and lateral
numerous arbitrations as a result. Although completely directions in relation to the geology.
weathered Grade V is an expected feature beneath the ● Seismic results in relation to results from other
Grade VI soil in tropical terrains, Grade V saprolite investigations, if available.
sometimes confusingly swaps places with the usually ● Conclusions and recommendations resulting from
deeper, and almost unjointed Grade I or II. (Saprolite the investigation that are of importance to the
is a weak, water sensitive, weathered in-place, some- project.
times beautifully structured and coloured relic of the
rock). Although reflection methods have eventually dom-
If this reversal of weathering grades appears in a tunnel inated the field of exploration seismics due to the various
arch beneath massive, high velocity core-stones, or if needs involved with deeper exploration, there is ‘univer-
there is a generally very undulating rock surface, with fre- sal’ use of shallow refraction seismic in sub-surface inves-
quent tunnel penetrations into weathered materials, tigations for civil engineering projects around the world,
there can be major delays. A tunnel collapse is difficult to due to its apparent simplicity and low cost. Further-
avoid when water is present, unless preparations have more, refraction seismics can be used to remove (from
been made, as a result of the more frequent exploratory the more deeply focussed reflection data), the ‘adverse’
drilling demanded when seismic anomalies such as these effect of the first meters or tens of meters of the hetero-
are suspected. geneous weathered layer, where differences in the ori-
Pre-injection ahead of the tunnel face, and heavier ginal rock quality may cause tens of meters of sub-surface
tunnel support, would be the very basic requirements ‘topography’ in the case of on-land exploration.
in a drill-and-blasted tunnel. (This is one of the purposes
of the ‘Q-system’ of rock mass characterization and
tunnel support selection). In the case of a TBM (tunnel 1.2 Some basic aspects concerning
boring machine) excavation, a change to a closed mode elastic body waves
in the case of a hybrid machine with earth-pressure-
balance (EPB) would be needed, especially if the wea- It is usually assumed that the strains associated with the
thered depressions in the bedrock contained water, as is passage of a seismic wave are of minute, sub-micron mag-
usually the case. nitude, and except in the neighbourhood of the source,
Best advice of all, as a direct result of a seismic refrac- the strains are generally assumed to be elastic. Based on
tion survey, would be to drive a deeper tunnel from the this assumption, the velocities of propagation of seismic
start. It is easy to imagine subway station construction waves are determined by the appropriate elastic moduli
under such heterogeneous conditions. It could be and densities of the materials passed through. The general
extremely time-consuming, and even dangerous. The form of the classic equations linking these three quan-
cost of deeper access to the stations, via longer escal- tities is V  (E/) ⁄ . Compressional bodywaves (primary
1
2

ators, would be a small price to pay for much reduced or P-waves) propagate by alternating compression and
tunnelling and station costs. dilation (Figure 1.1 a) in the direction of the waves.
Shallow seismic refraction, some basic theory, and the importance of rock type 5

The third important elastic modulus influencing the


conversion between dynamic properties is the bulk modu-
lus (K), defined as the ratio of the volumetric stress (P)
and the volumetric strain (v/v). Since the three mod-
uli are linked by the equation   K  4/3
, it
follows that Vp can also be expressed as:

1
 K  4 /3
 2

Vp    (1.3)
  

This equation therefore demonstrates the fundamen-


tally faster nature of Vp in relation to Vs. The ratio of these
two dynamic properties are also linked by the dynamic
Poisson’s ratio for the material, as will be shown in the
next section, which contains some standard equations.

1.2.1 Some sources of reduced


elastic moduli

In the case of micro-cracked, fractured, or jointed rock


masses, there is a correspondingly reduced set of moduli
in relation to the undisputed elastic nature of the intact
matrix, because of micro (and probably elastic) displace-
ments in normal and/or shear directions across and/
Figure 1.1 Elastic deformations and particle motions associated or along the micro-cracks, fractures or joints. These repre-
with the propagation of body waves: a) P-wave, sent an important part of the source of attenuation of
b) S-wave. Based on Bott, 1982. the seismic waves in the dry state, due both to various
scales of wave scattering and due to the intrinsic micro-
deformations. Added losses are incurred if these micro-
The oscillating uniaxial strain involved in the case of a or macro-discontinuities are partly saturated, since there
confined body, means that the axial modulus () con- is communication with the pores and eventual pore
trols the velocity of propagation, thus: fluid, and minute flows may be initiated to equilibrate
pressures. These micro-imbalances will only be equili-
1
 2 brated when the frequency is sufficiently low.
Vp    (1.1) The above mechanisms mean that dynamic proper-
  
ties, such as the velocities, Poisson’s ratio and attenuation
tend in practice to be dispersive, or frequency depend-
Shear bodywave waves, termed secondary, transverse ent. They are also of course rock quality and environment-
or S-waves propagate by a sinusoidal pure shear strain dependent, in the broadest possible meanings of these
(Figure 1.1 b) in a direction perpendicular to the direc- words. As rock quality declines, or the surface is
tion of the waves. The shear modulus (
), which is approached, there develops a serious discrepancy between
given by the ratio of shear stress ( ) divided by the shear the dynamic or elastic properties of the intact matrix
strain (tan ), will therefore control the (lower) velocity and the dynamic properties of the (partly discontinu-
of propagation, thus: ous) medium. The ratio between the dynamic proper-
ties of the (partly discontinuous) medium and the static


1
2 deformation properties, such as the (rock mechanics)
Vs    (1.2) deformation moduli and joint stiffnesses (the inverse
  
of compliances), may rise into double figures in this
6 Rock quality, seismic velocity, attenuation and anisotropy

complex region, where velocity-depth gradients are


often extreme.
At depth, under high confinement, and if rock quality
is high, it is assumed that there will be only small dis-
crepancies between the dynamic properties of the matrix
and the dynamic properties of the rock mass. The ‘static’
crack and joint stiffnesses, being so high, will be close to
the dynamic (inverted) crack and joint compliances.
There is controversy however, about the ratio of the
dynamic normal and shear compliances, and the
(inversed) ratio of the ‘static’ normal and shear stiffnesses.
There is even controversy over whether friction is a valid
attenuation mechanism, at the level of these micro-
displacements.
In the rock mechanics of ‘static’, ‘macro-deformations’,
we are familiar with a significant mismatch between the
high normal stiffness, and the much lower (and scale-
dependent) shear stiffness. Concerning the ratio of
dynamic compliances, geophysicists seem not to be so
sure, a dilemma that also probably affects whether fric-
tion is, after all, to be a valid attenuation mechanism, as
assumed in much of the geophysics literature, virtually
up to the present day.
Figure 1.2 Mean Vp and Vs statistics from 4 km of seismic profiles
in metamorphic and igneous rocks. Sjøgren, 1984.
1.3 Relationships between Vp and
Vs and their meaning in
field work
The rock mass quality Q-value, mentioned earlier in
the introduction, is composed of parameter pairs (RQD/
The advantages of using both P-wave and S-wave data to
Jn, Jr/Ja and Jw/SRF – see Appendix A for descriptions
interpret seismic results in hard rock was strongly empha-
and ratings). These effectively describe relative block size,
sised by Sjøgren, 1984. This has been reinforced by the
the inter-block friction coefficient, and an active stress. This
successively easier acquisition of multi-component, multi-
rock quality term will be utilised in various places in this
channel data, and rapidly developing PC analysis cap-
book, not least as a possible interpretation of seismic
abilities. In addition to many other sets of data, some of
quality Q (the inverse of attenuation). (Improvements
which will be referred to later, Sjøgren, 1984 presented
to equation 1.4 will be developed in Chapter 5, to allow
average Vp/Vs ratios from 93 rock sections from 5 differ-
for its application to weaker and more porous rock
ent sites in igneous and metamorphic rocks. These are
types, and to adjust it for depth or stress effects).
reproduced in Figure 1.2.
A rock quality Q-value of 1 tends to be heavily
P-wave velocities ranged from 3.3 to 5.7 km/s, and
jointed, containing some clay, while values 1, are tend-
S-wave velocities from 1.6 to 3.4 km/s. On average, Vp/Vs
ing towards better quality, with wider spacing of joints,
ratios were 1.89 in the rock mass with lower velocities
less joint sets, and no clay. (Q may reach values of about
(heavier jointing) and 1.80 in the rock mass with higher
1000 to 2000 in the case of massive, unjointed rock
velocities (sparser jointing).
masses, confined at depths of say 500 m or more).
These two ratios of Vp/Vs imply rock mass quality
The ratio Vp/Vs depends on dynamic Poisson’s ratio
Q-values of roughly 1 to 10, and 10 to 100 respectively,
() according to the following:
according to the following near-surface, hard rock Vp–Q
relationship (Barton,1991).
Vp 2  2
 (1.5)
Vp ≈ 3.5  log10 Q (1.4) Vs 1  2
Shallow seismic refraction, some basic theory, and the importance of rock type 7

From equation 1.5 one can derive the value of dynamic In the case of massive rocks of low porosity, the static
Poisson’s ratio as follows: and dynamic values of the elastic constants (e.g., the
elastic moduli Estat. and Edyn.) are quite close, while for
( Vp / Vs )2  2 heavily fractured and clay bearing zones, large differ-
 (1.6) ences between Estat and Edyn are seen (e.g. Cosma, 1995).
2( Vp / Vs )2  1
Rock mass quality descriptions such as Q or RMR or
RQD, which are described in more detail later, correlate

The ratio Vp/Vs is about 3 for hard (zero-porosity) better with static moduli than with dynamic moduli.
rocks, for which   0.25. However, in the case of Numerous relations between these moduli will be given
unconsolidated sediments, the ratio Vp/Vs can even reach in Chapter 6.
values of 20 to 40 for near surface material, for which  is
commonly greater than 0.45. Later in this chapter, high
values of (dynamic) Poisson’s ratio for a near-surface fault 1.4 Some advantages of shear
zone will also be seen, for similar reasons to the above. waves
A rock quality interpretation, linking these dynamic
parameters, can be added here, by taking Sjøgren’s (1984) In addressing the challenge of resolving the 0–50 m reso-
P- and S-wave results from 4.1 km of seismic profiles lution problem, Dasios et al., 1999 reported multi-
for hard but sometimes weathered metamorphic and component investigations at four shallow sites (thick
igneous rocks (Figure 1.2). The mean value of Vp/Vs  clays, clay/sand sequences over chalk, mudstone overly-
1.89 in the more heavily jointed rocks (perhaps a rock ing granodiorite bedrock, and landfill), using a combin-
quality Q  1–10), and the mean value of Vp/Vs 1.80 ation of both compressional and shear wave seismic.
in sparsely jointed rocks (perhaps a rock quality Q  The authors of course admit that there is a higher
10–100), can be used to calculate dynamic Poisson’s ratios level of effort required to conduct multi-component
of 0.30 and 0.28 respectively. seismic, requiring a three-component source configur-
As lower rock quality Q-values are approached in shear ation, and three-component geophones, but otherwise
zones and faulted zones (e.g. rock quality Q  0.1), the conventional multi-channel seismic recording systems,
ratio Vp/Vs increases to about 2.0, corresponding to a and PC-based processing software. Obviously the sur-
calculated value of dynamic Poisson’s ratio of about 0.33. veys are more difficult and more time consuming than
The corresponding Q-value (from equation 1.4, using compression-wave refraction or reflection, but the level
minimum Vp data from Figure 1.2) is indeed about 0.1. of geophysical information is that much more useful.
Extremely low Q-values, for example Q  0.01–0.001 They varied the acquisition geometry to optimize
(when Vp  1.5–2.5 km/s) will be needed before results. They found that under all the conditions, shear-
dynamic Poisson’s ratio values become as large as 0.45 waves penetrated with less attenuation than compres-
(as indicated for near-surface shear zones, in a later sional waves, also being unaffected by water saturation.
section of this chapter). Shear-wave reflections from shallow interfaces were in
Further basic equations linking Vp, Vs, dynamic some cases less affected by noise compared with the
Poisson’s ratio (), density () and dynamic Young’s modu- equivalent compressional-wave reflections.
lus Edyn. are as follows (Darracott and Orr, 1976): They offered the following simple explanation of
why shear-waves offer better vertical resolution than
 Edyn.(1   ) 
1
2 compressional-waves, particularly in shallow, unconsoli-
Vp   

(1.7) dated sediments. The dominant reason is that the
 (1   )(1  2 )  shear-wave velocities in such cases, are only a fraction
(sometimes less than one fifth) of the compressional-wave
1 velocities. This results in very small wavelengths, despite
 Edyn.  2

Vs   

(1.8) the fact that the dominant frequency of shear wave data
   2(1   )  is generally lower than is the case for compressional wave
data. In order to obtain the same level of resolution with
P-waves, energy of very high dominant frequency has to
3( Vp / Vs )2  4 (1.9) be generated, and this is correspondingly more attenuated
Edyn.  Vs2   
( Vp / Vs )2  1 in the low seismic Q sub-surface.
8 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(b)

Figure 1.3 a) Shear-wave velocities (km/s) and Vp/Vs ratios versus depth. b) Shear-wave velocities (km/s) and dynamic Poisson’s ratio versus
depth for a clay-over-mudstone-above-basement sequence, with an interpreted water table at 4.7 m depth. Results of multi-
component seismic at one of four shallow sites described by Dasios et al., 1999.

Since shear-waves are not attenuated at the water table, of Vp, Vs and the dynamic Poisson’s ratio, in the presence
and are little affected by changes in fluid saturation, they of a water table, and also show the effect of increasing
can more easily detect lithological changes with corres- depth in a uniform sedimentary rock. At their site, thin
pondingly less ambiguous velocity contrasts. The authors clays overlayed mudstone, with a basement of granodi-
found that under conditions of full water saturation, orite at 70 m depth.
P-wave velocity contrasts between lithologies were small, Figure 1.3 shows multi-component plots of shear-wave
whereas the shear-wave velocities reflected the true litho- velocity versus the ratio of Vp/Vs, and of shear-wave
logical changes. velocity versus (dynamic) Poisson’s ratio, each as a func-
In this brief summary, the results from a site they inves- tion of depth to the basement rock at 70 m depth.
tigated in Crewekerne, Dorset will be reproduced. Their Analysis of the P-wave first arrivals gave velocities of
results give a good illustration of the subtle interaction 496 m/s and 1,766 m/s for the unsaturated and saturated
Shallow seismic refraction, some basic theory, and the importance of rock type 9

layers, and indicated a depth of 4.7 m for the top of the 1.5 Basic estimation of rock-type
saturated zone. The dynamic Poisson’s ratio showed a and rock mass condition, from
small decrease close to the surface as a result of the con- shallow seismic P-wave velocity
solidation, then a sharp increase as a result of the water
saturation, followed by a steady decrease with depth When first investigating the bedrock for suitability for
within the uniform water-saturated mudstones. near-surface tunnelling or other relatively shallow con-
As expected from the theoretical calculation of the struction in rock, the preliminary use of shallow refrac-
dynamic Poisson’s ratio, there is a certain accentuation tion seismic is very typical, where surface access (including
of the above trends for the case of the ratio Vp/Vs, except noise) do not present major problems. As a minimum,
that this ratio reduces faster at shallow depth (rather the information gives a Vp – depth profile of inestimable
than when deeper), due to the strong Vp gradient. The value for further planning of the sub-surface investiga-
authors point out that the P-wave velocity was not avail- tion, in particular the optimal siting of boreholes for
able at the greatest depths, due to attenuation, so the future core-logging and permeability testing.
data was extrapolated to 70 m. There were indications of Figure 1.4 reproduces four examples of shallow refrac-
shear-wave anisotropy in the uppermost meters of clay, tion results from Sjøgren 1984, demonstrating the help-
but whether due to desication fractures or some form of ful information about the location, width and depth of
layering is not certain. zones of lower velocity. Later in this chapter, and in sub-
Although outside the usual range of Vp/Vs (about 1.4 sequent chapters, we will be seeing the many ways of
to 2.0) for water saturated rocks, it is of interest to see interpreting such velocities in terms of rock quality and
details of the development of Vp/Vs ratios in unconsoli- degree of fracturing, each tempered by the effect of rock
dated (e.g., subsea) sediments, by noting the progres- type, density, porosity, depth (or stress level), and of
sion from soft soils through compacted soils, to rocks. course the possible anisotropy (or directional depend-
Hamilton, 1979 gave comprehensive Vp-depth, Vs-depth ence) of the result in relation to an anisotropic jointing
and Vp/Vs-depth data for silty clays, turbidites and mud- frequency, and horizontal stress anisotropy.
stones to 1 km depth. Vp values increased slowly from The seismic refraction survey provides numerous
about 1.5 to 2.3 km/s as porosity reduced with increased depth to bedrock and quality of bedrock assessments at
depth, while Vs increased rapidly from only .05 km/s a small fraction of the cost and time needed for drilling.
close to the surface, to 0.15 km/s at depth. Vp/Vs ratios Depths are given at the impact points (hammer or shal-
therefore reduced very rapidly from double figures down low explosive source) and at the detector points (geo-
to about 2.5 at 1000 metres depth. The dynamic values phones or 3D seismometers), so a close spacing of
of Poisson’s ratio decreased, as a result, from about 0.49 detectors gives the equivalent of a large number of sound-
near the surface to 0.41 at 1000 m depth. ings or borings.
In connection with these high values of Poisson’s Sjøgren, 1984, gives the example of 5 m detector and
ratio for sediments, it is significant to note the relatively 25 m source separations for a 10 m deep bedrock inves-
high values of (dynamic) Poisson’s ratio that tend to be tigation. A 100 m profile gives the equivalent of 250 m
recorded in shear zones and fault zones at much shal- of soundings, and a complete distribution of relative
lower rock engineering projects. Gardener, 1992, calcu- quality beneath the profile. With the 10 m source and
lated the values of Poisson’s ratio from Vp and Vs 50 m detector separations needed for a deeper survey to
measurements at the Transfynydd power station in Wales, 50 m depth, the equivalent of 650 m of soundings per
estimating 0.45 for the shear zones, where the Vp velocity 100 m profile is given.
range was 1.6–2.7 km/s. The knowledge and experience of the geophysical team
The higher values of Poisson’s ratio for shear zones is essential in setting out optimal profiles in relation to the
have pseudo-static parallels to the special feature of geology and structural geology, in particular in relation to
heavily jointed rock masses, which can show ‘expansion anisotropic, layered media, and in relation to fault and
ratios’ or pseudo-Poisson’s ratios far in excess of 0.5, and shear zones. ‘Correct’ interpretation of the calculated
even in excess of 1.0 as (shear) failure is approached information cannot be divorced from the geology, since a
(Barton and Bandis, 1982). Elastic continuum theory given velocity (Vp or Vs or dynamic Poisson’s ratio) is not
is of course ‘violated’ by the shear displacements tend- unique to any one material but part of a scale or gradation
ing to occur on the failing joint surfaces. in the specific geological profile at the site, and reflects
10 Rock quality, seismic velocity, attenuation and anisotropy

Figure 1.4 Seismic refraction results illustrating the wealth of potential information obtained concerning near surface conditions. Sjøgren, 1984.
Shallow seismic refraction, some basic theory, and the importance of rock type 11

Figure 1.5 Typical ranges of Vp for sediments and for little weathered, moderately fractured rocks. Sjøgren, 1984.

various ‘environmental’ factors acting on each rock A similar range of values from the SSDS Project
domain, as will be demonstrated in subsequent chapters. granites in Hong Kong (Gardener, 1992) gives a useful
The later geological and rock quality interpretation qualitative impression of variations caused by weather-
of core recovered from boreholes drilled close to the ing and jointing in the same rock type.
seismic profiles is the domain of engineering geologists,
who besides identifying rock type, will perform careful Table 1.2 Typical range of Vp (km/s) for granite (Gardener, 1992).
logging of RQD, joint or fracture spacing, joint rough- Decomposed granite (soil) 1.6–1.8
ness and discontinuity mineral filling identification (or Fracture zones 2.8–3.5
testing). The performance of rock quality characteriza- Jointed granite 3.5–4.5
tion of drillcore is also standard practice for civil engin- Intact granite 4.5–6.5
eering and many mining projects, using the Q-value
(Barton et al., 1974, Barton, 2002) and RMR
(Bieniawski, 1989) as a minimum. Although these two At the hazardous second Severn Estuary crossing
methods have similarities, and common goals, there are between England and Wales, tidal currents are so strong
differences, and care is needed in converting Q to RMR that 85% of the crossing had continuous rock outcrops
and visa versa, e.g. Barton, 1995. between low and high tide. Sonar buoys and bottom
Typical ranges of velocities for relatively competent (lit- drag cable gave the following relatively tight ranges of
tle weathered moderately fractured) rocks are given in velocities for five rock types that were confirmed with
Figure 1.5. Much lower velocities, covering most of the boreholes, enabling the rocks to be identified across
lower diagonal space between 1 km/s and 6 km/s are seen the site.
with extremes of weathering, jointing and fault related
fracturing. The following is an example of the effects of Table 1.3 P-wave velocities at the Second Severn Crossing
weathering for just one rock type, from Sjøgren, 1984: (Gardener, 1992).

Average velocity Velocity range


Rock description (km/s) (km/s)
Table 1.1 Typical range of Vp for gneiss (Sjøgren, 1984).
Triassic mudstone 2.1 1.7–2.3
500 m/s Soil (above water table) Triassic siltstone 2.4 2.2–2.6
1700 m/s Highly weathered biotite gneiss Triassic sandstone 2.6 1.8–3.1
2800 m/s Weathered biotite gneiss Carboniferous siltstone
3500 m/s Jointed biotite/granitic gneiss and sandstone 3.0 2.5–4.4
4900–5400 m/s Sound biotite gneiss Carboniferous sandstone 4.0 3.0–4.4
12 Rock quality, seismic velocity, attenuation and anisotropy

Table 1.4 A list of typical Vp and Vs values from Press, 1966.

Material Vp (m/s) Vs (m/s)

Alluvium 500–2100
Clay 1100–2500
Sand 200–2000
Glacial Till 400–1700
Sandstone 1400–4500
Shale, Slate 2300–4700
Limestone
Soft 1700–4200
Hard 2800–6400
Crystalline 5700–6400
Dolomite 3500–6900
Granite, Granodiorite 4600–6000 2800–3200
Diabase 5800–6000
Gabbro 6400–6700 3400–3600
Basalt 5400–6400 2700–3200
Schist 4200–4900 2500–3200
Figure 1.6 Typical ranges of Vp for common rock types. Griffiths Gneiss 3500–7500 3300–3700
and King, 1987. Water 1450
Air 335

Griffiths and King, 1987, also give typical Vp ranges Table 1.5 P-wave velocities for different horizons in a river bank
for common rock types. These are reproduced in Figure terrace (Sen and Bandyoadhyay, 1990).
1.6, as a source of cross-referencing. Fractured, faulted 0.7–0.8 km/s Clayey and silty soil with pebbles
and heavily jointed zones extend the six major ranges 1.7–1.8 km/s Pebbles/cobbles in silty sand matrix
for these rocks far to the left on occasion. Note the 2.1–2.2 km/s Cobbles/boulders in silty sand matrix
extremely high velocities of the dense, ultramafic rocks,
which lie outside the common range of 1 to 6 km/s.
Table 1.6 P-wave velocities for phyllites at a Himalayan dam site
A comprehensive set of in situ seismic Vp values, and
(Dhawan et al., 1983).
some Vs values for common rock types, is also shown in
Table 1.4. The data are given by Press, 1966. Overburden/weathered phyllites 925–1200
The wide ranges of velocity for sandstone, shale, lime- Unweathered phyllites 2520–4500
stone and dolomite are mainly due to the wide ranges of
porosity (and density) for these materials. The surpris-
authors also give data for phyllites, which do not appear
ingly high range for gneiss is due to the wide range of
on the foregoing figures or tables of Vp data.
mineralogical composition (and density) for this rock.
The marked variation of velocities that are measured
in superficial deposits (0.5 to 2.0 km/s in Figure 1.5) 1.6 Some preliminary conversions
are partly caused by location either above or below the from velocity to rock quality
water table, as shown by Sjøgren’s 1984 data set. The
list given in Table 1.4 shows 0.2 to 2.0 km/s just for Due to the seismic ‘visibility’ of jointing in the upper 25
the case of sand, mostly for this reason. to 30 metres or so, Sjøgren et al., 1979 and Sjøgren, 1984
The following is perhaps a good example of the influ- and others, have been able to record significant correl-
ence of particle size in river born sediments. ations between Vp, RQD and joint frequency. These
The last line of Table 1.5 (for cobbles and boulders) authors compared the results from a total of 113 km of
differs significantly from the range 1.3 to 1.9 km/s for P-wave profiles from fifteen sites, with the results of
‘river boulders’ given by Dhawan et al., 1983, presum- 2.85 km of core-logging from seventy four drill holes at
ably due to differences in porosities. In the latter case, eight of the hard rock sites. (They also had 5 km of
the ‘silty sand matrix’ is presumably absent. These last S-wave surveys at five of the sites).
Shallow seismic refraction, some basic theory, and the importance of rock type 13

The range of rocks occurring at the measured locations, Figure 1.7 reproduces Sjøgren et al., 1979 data in the
mostly in Norway, were: amphibolite, granite, gneiss, readily absorbed format used by the authors. Mean values
meta-anorthosite, pegmatite, porphyry, quartzite and for all the parameters apply. The Edyn modulus is the
mylonite. The authors were careful to emphasise that same as the symbol () used to denote the uniaxial or
the correlations they derived between P-wave velocity axial modulus in equation 1.1.
and jointing descriptions such as mean RQD and mean Since their measurements were shallow, the effect of
frequency F(m1) were relevant only to unweathered stress-induced joint closure was minimised. They also
igneous and metamorphic rocks, and generally for the effectively removed other sets of variables by generally
upper 20 to 30 metres. recording correlations for hard and almost unweathered
igneous and metamorphic rocks. The usual variables of
depth, porosity, uniaxial compressive strength and dens-
ity were therefore largely removed.
A hard rock, near surface correlation of Vp and the
rock quality Q-value was derived by Barton, 1995, on
the basis of trial and error fitting to cases known to the
writer and also Q-logged. (These will be described in
Chapter 5). In Figure 1.8, the important effects of poros-
ity, uniaxial strength and depth are ignored, as for the
Sjøgren et al., 1979 data. Note that the RQD and F/m
mean data have been smoothly extrapolated beyond
both ends of the Sjøgren et al., ‘data-base’ represented
by Figure 1.7.

1.7 Some limitations of the


refraction seismic velocity
interpretations
Figure 1.7 Mean values of physical and dynamic properties for
hard, unweathered igneous and metamorphic rocks,
The seismic refraction method has some important limi-
based on shallow refraction seismics. Sjøgren et al., 1979.
Q-Scale added by Barton, 1995.
tations. One is that the horizontal or sub-horizontal ray

Figure 1.8 RQD and Fm1 trends from Sjøgren et al., 1979. Q-scale, (as also in Figure 1.7) added by Barton, 1995. These results and
approximate correlations to seismic-frequency Vp are relevant for hard, low porosity, unweathered, near-surface rock masses.
14 Rock quality, seismic velocity, attenuation and anisotropy

paths record only the upper part of each seismic layer. A


thin high-speed layer can mask underlying material, while
a low-velocity intermediate layer will not be recognised
for similar reasons. Depth calculations to underlying
refractors will be erroneous. The hidden low velocity
zones can of course be detected by up-hole shooting from
a borehole to the seismic spread (i.e. VSP), or by inspec-
tion and index testing of core, if available.
A useful review of refraction seismic methods and com-
monly used methods of interpretation is given in Whitely,
1990. Citing Sjøgren, 1984 that low velocity zones are
consistently interpreted as being shallower than the bore-
hole confirmation, Whiteley, 1990 went on to compare
three interpretation methods of the simple yet frequently
occurring situation shown in Figure 1.9. The three solu-
tions were obtained by independent geophysics practi- Figure 1.9 Three independently derived depth estimates for a low
tioners. It is clear that control drilling of identified velocity zone detected by refraction seismics. Whiteley,
features must be made before making important deci- 1990.
sions such as minimum depth of cover over a sub-sea
tunnel.
Stapledon and Rissler, 1983 emphasised the follow-
ing potential shortcomings of the seismic method, as engineering projects may be far different from these
related to near-surface investigations for civil engineer- apparently uniform conditions. Yet even the ‘simple’ site
ing projects. may contain velocity anomalies, which reduce the image
quality. The time horizons suffer ‘push-down’ beneath
1. Minor geological defects such as weathered seams or slow-velocity anomalies, and ‘pull-up’ beneath fast
minor faults may govern the engineering behaviour velocity anomalies.
of a site, especially if their orientation is unfavour- The geological ‘reality’ shown on the left in Figure 1.10
able. The seismic method ‘generally is unable to contrasts nicely with the equivalent post-stack seismic
detect such defects’. Is this pessimism justified or is time section, shown on the right. There is significant
it entirely correct? distortion of layer ‘horizons’, and perhaps surprisingly
2. P-waves are first to arrive at the geophones ‘and must persistent follow-through to depth. Armstrong et al.,
take the shortest path through the best rock’. Will 2001 describe a method for compensating for these
this mean that they do not represent the average discrete overburden velocity anomalies. (Their warning
local quality but the best local quality? about possible miss-calculation of reserves, or miss-
positioning of infill wells, or by-passing of incremental
Both shallow and deep refraction seismic are subject reserves in relation to petroleum engineering, can clearly
to velocity anomalies, causing time-distortion and there- apply in other contexts to shallow civil engineering
fore depth anomalies, as we have seen in the above shal- projects, where for example there are buried sediment-
low example. Although the following case of Armstrong filled channels that have been known to plague certain
et al., 2001 is strictly speaking a deeper reservoir case, it tunnel projects, due to high water storage, and
provides such a nice illustration of the hazards of depth ‘constant’ inflow pressure).
interpretation that it will be referred to in this first Also concerning the possibility of depth estimation
chapter. error, it was emphasised by Bradford and Sawyer, 2002,
As the authors point out, the overburden above a in the context of shallow seismic reflection measurements,
hydrocarbon prospect is often more or less horizontal or that larger depth and layer thickness estimates would
perhaps with gently dipping sedimentary layers. It is gen- occur when using conventional velocity analysis, if in the
erally paid little attention, in relation to the focussed presence of the extreme velocity gradient close to the dry-
investigation of the reservoir target at depth. Clearly, saturated transition, (usually fairly close to the surface in
the overburden at typical shallow refraction sites for civil temperate climates).
Shallow seismic refraction, some basic theory, and the importance of rock type 15

(a) (b)

Figure 1.10 Schematic cross-section through an overburden containing velocity anomalies of different geological age. The idealized vertical
section is shown on the left, and the contrasting, time-horizon distortions, post stacking, are shown on the right. Armstrong,
et al., 2001.

The authors indicated (as also seen earlier), that the Sjøgren, 2000 expressed concern about the useful-
P-wave velocity could increase by a factor of four or ness of the GRM method for near-surface geotechnical
more at this transition, in the case of unconsolidated investigations, where details of the various overburden
sediments changing from dry, through partial satur- layers are required since they may have important con-
ation, to full saturation – which may occur just above sequences for the subsequent geotechnical design.
the water table, due to the action of capillary forces. A A relatively more recent technique for modelling of
large velocity gradient (e.g. 400 m/s to 1600 m/s from 1 travel times and travel time inversion in refraction seis-
to 10 m depth) apparently violates many of the assump- mics is the so-called Eikonal solver. In principle, this
tions made in conventional reflection data processing involves the calculation of travel times on a regular
schemes. velocity grid. Early versions, originating from the late
In a recent paper, Sjøgren, 2000, evaluated several 1980s were restricted to a plane topography for the
standard methods for interpreting travel time curves. recording surface.
He utilised the ABC method (originating from 1931), Lecomte et al., 2000, describe a first order Eikonal
the ABEM correction method (early 1950s, detailed in solver that can incorporate the exact topography of the
Sjøgren, 1984), the mean-minus-T method (mid-fifties, surface terrain, and any arbitrary lateral variation of
also adopted in the ABEM method), and the Hales velocity. There is no restriction on the velocity contrast.
method (1958), in order to critically evaluate a more In effect, the model is built up layer by layer, with the
recent (1980) generalized reciprocal method (GRM) of refractor imaging, and the velocity mapping being per-
Palmer, in particular Palmer, 1991. formed for each identified refractor at a time, as seen
16 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(b)

(c)

Figure 1.11 An arbitrarily chosen model for demonstrating a new method of refraction seismic inversion. Lecomte et al., 2000 used this
synthetic model for demonstrating some basic elements of the Eikonal solver, which calculates the travel time of the fastest
wave at any point of a regular velocity grid, using the head waves generated in refraction seismics. The three stages of model-
ling shown here are: a) Wave-fronts and raypaths at receivers along the surface, when considering the whole model (but
minus the acoustic wave velocity in the air, which was omitted for clarity. b) Wave-fronts and raypaths at receivers along the sur-
face, when masking the deepest layer. c) Wave-fronts and raypaths at receivers along the surface, when masking the two deepest
layers.

schematically in one of their illustrative figures, repro- Consider for example the jointed chalk marl at the
duced here in Figure 1.11. Chinnor Tunnel in the UK, where jointing in this weak
material were seismically closing at about 15 metres depth,
to give a stable 1.6 km/s field velocity, despite changes in
1.8 Assumed limitations may the degree of jointing (‘rock’ Hudson et al., 1980). This
hide the strengths of the can be contrasted with the jointed gneiss at the Gjøvik
method cavern in Norway, which gave a continuous rise in vel-
ocity from 3.5 to 5.5 km/s in the first 50 metres depth due
In later chapters the numerous factors influencing seismic to increased stress, yet had almost unchanged rock qual-
velocities such as joint frequency, porosity, rock (and joint ity. The joint frequency, RQD and rock quality Q-values
wall) strength, density, depth, stress, stress anisotropy, did not show an improvement with depth (Barton et al.,
degree of saturation, type of saturating fluid, will each be 1994). Both the above observations could be interpreted
reviewed. This will be done in order to emphasise that as ‘limitations’ of the method. In fact they are demonstrat-
the seismic method has numerous complications, but also ing specific and quite logical physical laws of behaviour.
inestimable advantages, and that some of the assumed The latter is an example of the need to interpret seismic
shortcomings can be due to misinterpretations (often an velocity with knowledge of depth and/or stress level, since
over-simplification, or perhaps even a pre-conceived a rock quality Q-value increase from perhaps 1 to 100
opinion about ‘seismic’ limitations). might otherwise be assumed in these first 50 metres,
Shallow seismic refraction, some basic theory, and the importance of rock type 17

based on a ‘hard rock’, ‘stress-less’ interpretation (i.e. round of advance would also be demonstrating a
equation 1.4, and Figure 1.8). ‘reversed’ Vp-quality behaviour. However, the more dom-
The interpretation cannot be divorced from consider- inant effect of a reduced need for time consuming rock
ation of what is actually occurring in a rock mass as depth support would be consistent with the normal ‘high
increases, i.e. joint normal stiffness increases, joint aper- velocity  high quality’ concept. Since it is more logical
ture reduces, joint frequency reduces (usually), and the that Vp increases should accompany rock quality or
clay content in the joints reduces (usually). The list can Q-value increases, such cases of ‘reversed’ quality must
be lengthened considerably by adding that the stress be treated as separate, method-specific cases.
increases, the deformation modulus increases, the permea-
bility reduces, the pore pressure increases. Since there
may be changes of lithology with depth increase, the 1.9 Seismic quality Q and apparent
provision ‘usually’ should really be added to each of similarities to Q-rock
the above.
When considering the possibility of relating seismic A fundamental feature of the propagation of stress waves
velocity to a rock quality descriptor such as the rock in all materials is the absorption of energy and the result-
quality Q-value, another quite basic problem must ing change in shape of the transient waves. Non-linear
also be considered. A velocity of 2.5 km/s for massive friction has traditionally been assumed to be one of the
chalk marl of high porosity, as in the better parts of dominant attenuation mechanisms (Kjartansson, 1979),
the UK-France Channel Tunnel, will have entirely dif- but as will be discussed in Part II, there are a large num-
ferent engineering consequences to that of a regional ber of mechanisms that can explain the different degrees
fault-zone of the same 2.5 km/s velocity, crossing a of attenuation in different rock masses, including scat-
Japanese high-speed rail tunnel excavation, and delay- tering due to structure, and squirt flow.
ing progress by months. In this best quality chalk marl, Geophysicists commonly characterise seismic attenu-
near-world record speeds of (TBM) tunnel boring were ation by the seismic quality factor Q, which of necessity
achieved. has been termed Q seis throughout this book, to distin-
The natural velocity of the unjointed rock under in guish it from the rock mass quality Q. Intuitively Q seis
situ conditions (Sjøgren et al., 1979), and the contrast is related to, but no relation of, the rock mass quality Q
seen in low velocity zones is the main index of difficulty, of Barton et al., 1974, which is a ‘quality number’ also
since an order of magnitude reduction in the rock quality widely used in civil and mining engineering in the last
Q-value will generally accompany each 1.0 km/s reduc- several decades.
tion in seismic velocity, according to the simplified The seismic quality Q seis is often defined as the max-
model shown in Figure 1.8. imum energy (Emax) stored during a cycle, divided by
Low velocity and potentially high permeability zones the energy lost (E) during a cycle:
will be the natural focus of attention in most sub-surface
civil engineering projects. However, in a TBM (machine- 2 (1.10)
Q seis   E max
bored) tunnelling project, there will be serious delays if E
there is too much high velocity rock, due to the slow
progress made in hard, sparsely jointed rock. This would For dry rocks, Q seis has been claimed to be independ-
give a ‘reversed’ Vp–rock quality indication, due specif- ent of frequency over a reasonable frequency range
ically to poor borability and the need for frequent (McKenzie et al., 1982). However, as will be seen in
cutter changes. Part II, Q seis is a remarkably sensitive indicator of
A Q-value based ‘Qtbm’ rock-machine quality factor anisotropy, and is frequency dependent in the case of
has been developed for this specific problem (Barton, saturated or partly saturated conditions. When attenu-
2000), which also allows for the fact that more jointing ation (Q1seis) is high, Q seis obviously has a low value.
is good for progress – up to some limit, when other prob- Chapter 10 in Part II, will address Q seis in detail, and
lems may arise. (More tunnel support is needed, there also draw some tentative parallels between ‘Q and Q’,
could be gripper-setting problems, the cutter head could via a common link to deformation properties.
even be jammed). Intuitively, the rock mass quality Q (of Barton et al.,
In the case of drill-and-blasted tunnels, the drillability 1974) which has a high value in high-modulus, high-
and blastability components of the cycle time for one velocity rock masses (i.e., Q  100, Vp  5.5 km/s,
18 Rock quality, seismic velocity, attenuation and anisotropy

Emass  45 GPa) would seem likely to have high energy Further discussion of Q seis is given in Chapter 4, based
storage and low energy loss under such conditions, and on recent seismic attenuation tomography (see alterna-
therefore qualify for a high value of Q seis and a low tive tomograms). However, a much fuller treatment of
value of attenuation (Q1 seis). Conversely, in a rock mass Q seis is given of both lab-scale and in situ scale attenuation
characterized by many joint sets with clay coatings and phenomena, in Part II, chapter 10, and in relevant parts
fillings on many joints (i.e., Q  0.1, Vp  2.5 km/s, of Chapters 13 and 15.
Emass  5 GPa), both low energy storage and high
energy loss per cycle would be expected. (Q seis is low
and attenuation Q1seis is correspondingly high.)
2 Environmental effects on
velocity

In this chapter, the effects of near-surface weathering on and density  in gm/cm3. Seismic velocities ranged from
the seismic velocity Vp will be reviewed. This automat- 2.3 to 6.5 km/s, and densities from 2.1 to 3.0 gm/cm3.
ically introduces the separate, but closely integrated Early Bulgarian experiences with seismic registration
effects of density, porosity, uniaxial compressive strength, of weathering effects are provided by Iliev, 1966. Fresh
and the depth and degree of saturation. Depth and stress and weathered monzonite were shown to have the fol-
effects will only be superficially reviewed here; that dis- lowing ranges of properties, and linear relationships
cussion belongs in later chapters of Part I dealing with between Vp–n% and Vp– (gm/cm3). (See Figure 2.2)
anisotropy and rock-burst or stress-slabbing in deep
excavations, and is of course a fundamental aspect of
all the deep or high pressure seismic results reviewed Vp km/s E (GPa)  (gm/cm3) n%
in Part II. The review will be loosely organised into sub- Fresh monzonite 5.0 50 2.61 2
sections, on density, porosity, uniaxial strength and Weathered monzonite 1.4 6 2.34 12
water content, but overlap will inevitably occur within
each sub-section. Weathering and depth effects are insep-
arable from the general presentation of reviewed data. The linear Vp–n% and Vp– relationships conceal a
The key result of the inter-relationships in this near- non-linear Vp–uniaxial compressive strength trend. When
surface environment, is a velocity-depth gradient even in a reduction in Vp due to c and due to n% are assumed,
one rock type, that can easily climb into two and three the strength of the porosity relationship as above, needs
figures, for example 2 km/s increase in the space of modification. Iliev, 1966, noted that a weathering coef-
20 m. The dual effects of rock mass strength and qual- ficient could usefully be defined as the ratio (Vpo–Vpw)/
ity improvement, and of vertical and horizontal stress Vpo where sub-scripts (o) and (w) signify fresh and
increase, are usually responsible. weathered. The coefficient approaches values of 0 and 1
at opposite ends of the weathering scale.
Many of the long span bridges in Japan have been
2.1 Density and VP constructed in soft rock such as Tertiary mudstones and
sandstones, or weathered Tertiary granites. The long
The strong influence of density on P-wave velocity, and span bridges of the Honshu - Shikoku Bridge system
the stabilisation of density below the weathered zone are described by Yamamoto et al., 1995, and Ishikawa et al.,
nicely demonstrated in Figure 2.1 (Ikeda, 1993). The 1995, had foundation sizes in the 50 to 100 m range but
marked fluctuations in velocity at depth were interpreted nevertheless had contact pressures as high as 1 to 2 MPa.
by the author as due to high shear stresses, which were For this reason, Japanese authorities devised compre-
interpreted from hydraulic fracturing conducted in the hensive routines for geological and geotechnical investi-
same holes. (Presumably the elastic isotropic estimate of gations. Seismic methods, in situ deformability, strength
H max, based on the ‘3P-Q’ model was several times the and classification schemes were used extensively, espe-
magnitude of the measured H min). Note the typical cially when trying to extrapolate the results of in situ
rapid increase in velocities in the 25 to 75 m depth zone, testing at more convenient onshore sites, to the actual
which partly mirror density increases and are partly undersea locations of the pier foundations.
related to joint closure and less frequent jointing. The data given in Figure 2.3 shows in situ seismic-
In the case of a range of rock types including marl and velocity-based rock classes, porosities, densities and
peridotite, Kujundzíc and Grujíc, 1966 found a linear degree of saturation, each intimately linked. This remark-
relation Vp  4.75 – 7.3 (r  0.88) for Vp in km/s, able, and useful diagram covers all the sub-titles of this
20 Rock quality, seismic velocity, attenuation and anisotropy

chapter, but has been included at the beginning, together seismic rock class versus shear strength, and deform-
with density, to indicate the integrated nature of these ation modulus versus RQD and uniaxial strength. Most
measures of rock mass quality (or lack of quality). Other of the data was obtained from measurements in
useful data sets include Vp versus deformation modulus, medium to lower quality, weathered granites.

Figure 2.1 Influences of weathering, depth of measurement and density on Vp and resistivity. Ikeda, 1993.
Environmental effects on velocity 21

The authors used an extended version of the Tanaka classes at the lowest end of the scale (DH, DM, DL) for
and Japan Highways classification (which involved the rock masses with velocities in the range 1.5 to 2.7 km/s.
six classes A, B, CH, CM, CL and D where subscripts Table 2.1 shows the scheme adopted, which cross-cor-
mean high, medium and low), and included three relates with deformation modulus, density, porosity
and resistivity.
A large collection of laboratory Vp– (gm/cm3), and
Vp–n% data is given by Kelsall et al., 1986, for the case
of basalts from California and dolerites from
S.W. England. Data that fall outside the general trend
for the intact rock are ascribed to fissured and persistently
microcracked rock, shown by the black data points in
Figures 2.4 and 2.5.
The lower seismic velocity of the fissured samples
is accentuated by the air-dried state of these samples.
When plotted on a log-linear scale, the uniaxial
Figure 2.2 Effects of weathering on Vp of monzonite are seen in strength is seen to broadly correlate with the air-dry
linear Vp-n% and Vp- relationships. Iliev, 1966. Vp value. The data set given in Figure 2.6 goes to

Figure 2.3 Inter-relationships between rock class, Vp, porosity, density and degree of saturation at the Honshu - Shikoku Bridge project in
Japan. Ishikawa et al., 1995.

Table 2.1 Extended (low velocity) Japanese classification scheme used at Honshu-Shikoku Bridges, showing cross-correlation of parameters.
Note extreme range of densities due to weathering. Yamamoto et al., 1995.

Rock class Vpr (km/s) Rt ( • m) Esb (GPa) c (103 kg/m3) R (103 kg/m3) nc (%) nR (%)

DL 1.5–1.8 1–4 0.05–0.3 1.7–2.1 1.55–1.75 33–58 50–64


DM 1.8–2.2 4–7 0.3–0.8 2.1–2.3 1.75–2.0 19–33 37–50
DH 2.2–2.7 7–12 0.8–1.5 2.3–2.5 2.0–2.15 11–19 27–37
CL 2.7–3.3 12–20 1.5–3.0 2.5–2.55 2.15–2.3 6–11 20–27
CM 3.3–4.0 20–50 3.0–6.0 2.55–2.6 2.3–2.4 3.5–6 15–20
CH 4.0–4.8 50–120 6.0–12.0 2.6–2.65 2.4–2.5 2.0–3.5 11–15

Note: Vpr: P-wave velocity of rock mass; Rt: resistivity of rock mass; Esb: deformation modulus from pressure meter; c: density of core; R:
density of rock mass; nc: porosity of core; nR: porosity of rock mass.
22 Rock quality, seismic velocity, attenuation and anisotropy

Figure 2.5 Effects of porosity variations on Vp for air-dry samples


of dolerites. Note increased density of the fissured sam-
ples, presumably indicating a change in composition.
Figure 2.4 Effects of dry density on Vp for air-dry samples of
Kelsall et al., 1986.
dolerites. Note increased density of fissured samples,
presumably indicating a subtle change in composition.
Kelsall et al., 1986.

of high stresses (1000 MPa or 10 kbars). The velocity of


a variety of high density rocks such as dunite and ser-
unusually high levels of strength (500 MPa) and veloc- pentinite are shown in Figure 2.8. For densities in the
ity (7.5 km/s); the latter a direct function of the high range of 2.5 to 4.5 gm/cc, velocities ranged from 6 to
density (2.9–3.1 gm/cm3) of the dolerites. Note the more than 9 km/s at these extremely high pressures
ratio of Vp (air-dry) to Vp (saturated) given in Figure (Birch, 1961).
2.7. The samples with the pre-existing fissures show Velocity (and of course density), have been used with
greatest contrast in these velocities, due to the positive success for identifying minerals from host rocks.
effect of wave transmission through water filled fissures Salisbury et al., 2000, used seismic imaging of known
(or joints). ore bodies in central and eastern Canada, together with
Before leaving this section on (mostly) Vp and density high pressure laboratory tests, using what they termed a
trends, caused by weathering, it is instructive to also look ‘crack closure pressure’ of 200 MPa confining pressure.
at extreme Vp values due to exceptionally high dens- They drew various envelopes in velocity-density space,
ities, both from natural causes and from the influence to distinguish commonly occurring sulphide ores, and
Environmental effects on velocity 23

Figure 2.7 Air-dry and saturated Vp values for intact and fissured
samples of dolerite. Kelsall et al., 1986.

Figure 2.6 Vp– c relation for high strength and weathered rocks.
Kelsall et al., 1986.

typical silicate host rocks. We may select some contrast-


ing combinations:

Pyrite: Vp  8.0 km/s density  5.0 gm/cm3


(the extreme member)
Pyrrhotite: Vp  4.7 km/s density  4.6 gm/cm3
Chalcopyrite: Vp  5.5 km/s density  4.1 gm/cm3 Figure 2.8 Extreme Vp-density data for crustal rocks at 1000 MPa
Serpentinite: Vp  5 to density  2.4 to confinement. Birch, 1961. (Numbers: next to open cir-
7 km/s 2.9 gm/cm3 cles  mean atomic weights; on diagonal lines  con-
(host rock) stant mean atomic weights (approx.)).

At the shallowest depths of the earth’s crust, namely elasticity-based, Hertz contact theory predicted that Vp
the soil cover, specific depth-density-Vp relationships should be proportional to the 1/6 power of the effective
are also evident. Brandt, 1955, developed a theory for stress. He then compared (in Figure 2.9) this Vp-depth
the influence of pressure and porosity (and saturation) gradient with test data for soil, clay and gravel meas-
on the seismic velocity in porous granular media. His ured by Nasu, 1940. The slopes of the test data plotted
24 Rock quality, seismic velocity, attenuation and anisotropy

Figure 2.10 Vp – porosity data for limestones, sandstones and


Figure 2.9 Effect of depth of burial on Vp values of soils in relation (jointed, weathered) granites. Fourmaintraux, 1975.
to theory. The gradient 1/6 is drawn from the origin,
signifying only the Vp-depth relation, not the velocity
magnitude, which can vary widely depending on the
composition of the soil. Brandt, 1955.
beyond extreme porosities of about 50%. Something
nearly approaching linearity is seen in the case of sand-
stones. A simple, illustrative set of experimental data,
on log (Vp)–log (depth) scales, ranged from 1/2 to 1/7,
applicable in civil engineering is that provided by Four-
bracketing his theoretical gradient prediction of 1/6.
maintraux, 1975, which is reproduced in Figure 2.10.
In practice this data and the accompanying theory can
Remarkable linearity is shown in these three cases.
help to explain the virtual seismic ‘disappearance’ of
The strong influence of the porosity of the matrix in
heavily stressed, faulted gouge at great depth or at large
rocks such as limestone and sandstone, and the linear
induced stress levels. Such was experienced, for example,
nature of the Vp–n% inverse relationship is clearly
ahead of a stuck TBM in an 800 m deep tunnel where
demonstrated. In the case of the granites, where joint
cross-hole tomography was designed to investigate a
porosity and presumably weathering of the matrix, are
known fault (Contract report, NGI, 1998).
the chief sources of porosity, the reduction of velocity is
even more marked. The uniaxial compressive strength is
also strongly related to matrix porosity in the case of
2.2 Porosity and VP porous rocks such as limestones and sandstones. It may
therefore be logical to allow for the influence of both c
There is a wealth of data in the literature concerning the and n% when seeking an integrated Vp– rock quality –
effect of the rock matrix porosity on the P-wave velocity. deformability chart, to be developed in subsequent
This is found in most abundance in the rock physics chapters.
investigations to do with petroleum reservoirs, and will Wilkens et al., 1984, found that the percentage of
be reviewed in Part II, Chapter 13, and elsewhere. clay content in sandstones had a marked effect on the
In general, an approximate inverse proportionality is P-wave velocity of dry samples. Although this particular
found between velocity and porosity, but there are many data set was related to petroleum reservoirs, and confin-
subtle variations bought about by, for example clay- ing pressure was consequently very high (50 MPa), the
content in sandstones. High pressure data (from Chapter data will be included in this chapter, as it gives a good
13) suggests a fairly strongly curved, concave relation illustration of the adverse influence of clay-content,
for porous (30 to 80%) marine sediments, with velocities which is a particularly relevant aspect of the near-surface,
falling rapidly at first, with a plateau of about 1.5 km/s due to weathering effects on some constituent minerals.
Environmental effects on velocity 25

Figure 2.11 Effect of % clay content on the variation of Vp for


dry sandstones, for given values of porosity. (Note: high
pressure data, for illustration of relative effects. See
Part II, Chapter 13 for high pressure rock physics
data). Wilkens et al., 1984.

Figure 2.12 c–Vp trends from Ohkubo and Teresaki, 1977, with
data from a tunnel site in basalt, tuffs and agglomerates.
Clearly the higher velocities given in the figure will not Won and Raper, 1997.
be so closely approached in the case of near-surface clay-
bearing sandstones, but the relative effects of clay-
content are illustrative. Evangelista and Pellegrino, 1990, referred to exten-
For a given porosity, say 20%, Vp was shown to range sive Japanese data assembled by Ogawa, 1986, in also
from 3.5 to 4.5 km/s due to clay content reducing from citing the potential link between porosity and uniaxial
15% to 5% (approx.). Figure 2.11 shows Vp–n% data compressive strength. Figure 2.13 shows the enormous
for dry sandstones, with the % of clay displayed next to influence that porosity has on uniaxial strength, indi-
the data. cating a bi-linear trend in a semi-log plot. The influence
of porosity on density, and the influence of uniaxial
strength on stiffness, means that several inter-related
physical properties play their role in increasing or decreas-
2.3 Uniaxial compressive strength ing seismic velocity. Microcracking, jointing, stress level
and VP and degree of saturation (including type of fluid) add to
the complexity, as will be extensively demonstrated in
Classification of uniaxial strength by means of seismic later chapters, both in Part I and Part II.
velocity alone is obviously suspect since porosity, density Several hundred uniaxial compression strength tests
and grain size will also be important to differing degrees. on flysch sandstones were conducted by Pininska, 1977,
However, if envelopes are used to separate the major rock in three orthogonal directions. The following general
groups, then c–Vp relationships become somewhat trend can be seen in their c–Vp plot:
clearer, as illustrated by Ohkubo and Teresaki, 1977,
and Won and Raper, 1997, in Figure 2.12. The open
circles are data for basalt, tuffs and agglomerates from Vp  2.0 3.0 4.0 5.0 5.5 (6.0) est. km/s
c  10 20 40 80 100 (160) est. (MPa)
investigations at a tunnel and highway cutting site in
Australia. Note the trend lines ( c  Vp3 and
c  0.25 Vp3) whose 4:1 range is still insufficient to However, the scatter of data was very large, and one
encompass the range of data produced in the Japanese could refer to ranges of the above velocities of as much
study (Ohkubo and Teresaki, 1977 Oyo Corporation. as 1.5 km/s in the enclosed region of the above tabula-
Technical Note RP-479). tion. The doubling of strength for each 1 km/s increase
26 Rock quality, seismic velocity, attenuation and anisotropy

Figure 2.13 The inter-relationship between porosity and uniaxial compressive strength. Ogawa, 1986 from Evangelista and Pellegrino,
1990.

Figure 2.14 Poor correlation of Vp,  and c is evident for shale, due in part, to the similar densities of component minerals. Lashkaripour
and Passaris, 1995.

in Vp is a good mean trend. We can add this to another Index tests such as the point load test and Schmidt
general trend, namely that the rock mass quality Q-value hammer test (with density included in the interpretation)
increases 10-fold for each 1 km/s increase in Vp for the are known to correlate reasonable closely with uniaxial
case of hard, low porosity rocks at shallow depth. compressive strength. For this reason, Wei and Liu, 1990,
Environmental effects on velocity 27

strength (1 MPa to 10 MPa) for 1 km/s increase in veloc-


ity (i.e., 1.5 km/s to 2.5 km/s). The ‘decade’ rule-of-
thumb, referred above, was again demonstrated for
much of the data. The Japanese data shows slightly lower
strengths (i.e., 5 MPa compared to 10 MPa) for a veloc-
ity of 2 km/s, though as with Pininska, 1977, the scatter
of data is large when porosity variations are not shown.

2.4 Weathering and moisture


content

Effects of weathering on the physical and seismic prop-


erties of four rock types from a dam site (quartz diorite)
and from three quarries (andesite, basalt and dacite)
were reported by Saito, 1981. This very comprehensive
study, involving hundreds of samples with different
weathering grades and porosities, gives a very useful
picture of some key trends between strength, hardness,
Figure 2.15 Vp- c data for Tertiary mudstones and sandstones from
porosity, degree of water saturation and P-wave veloc-
Japan. Aydan et al., 1992 and Sato et al., 1995.
ity. These behavioural trends are fundamental to an
understanding of the in situ behaviour, where the addi-
tion of joints to the cracks and pores tested here, adds
another layer of complexity.
used Vp, Schmidt rebound value (R) and point load test Saito, 1981, collected numerous block samples of the
(I50) to evaluate the weathering grade of four igneous different rocks and weathering grades and cast these in
rocks. Their database was very large (1069 I50 tests, 499 regular shaped concrete blocks, before coring cylindrical
Vp tests, and 1330 Schmidt hammer tests) and gave, samples for his tests. Schmidt (N-hammer) tests were
as one might expect, strong correlation to the pre- made on these larger blocks. Figure 2.16a illustrates and
determined weathering grades. describes the typical weathered zones (1 to 5), and an
In the case of shale, with its mixed content of similar idea of the ranges of compression strengths (dry sam-
density minerals, i.e. quartz (  2.66 gm/cm3), illite ples) and porosities are given in Figure 2.16b and
and montmorillonite (  2.61 gm/cm3), the correlation 2.16c. The very different porosities of the three volcanic
of compressive strength and density is inevitably poorer rocks compared to the crystalline quartz diorite are well
than for most rocks (Lashkaripour and Passaris, 1995). reflected in the clear separation of the Vp values shown
Since seismic velocity usually correlates well with dens- in Figure 2.16c.
ity, it is perhaps inevitable that in the case of shales, The extended Vp range of 1 km/s to almost 6 km/s
P-wave velocity is not a sensitive indicator of compressive was the result of the huge range of porosities (57% to
strength, as shown by the large spread of data in Figure 1%). When only uniaxial strength and velocity were
2.14 from the same authors. plotted (Figure 2.16d) the fundamental differences in
Water content on the other hand correlates extremely porosity were not seen due to the relatively high
closely with uniaxial strength for shales, e.g., c  strength of the three volcanic rock types. Figure 2.16e
90e(0.5w) (w  water content), based on the mean of shows how Saito’s Schmidt (N) hammer rebound data
data from two coalmines. The porosity and com- related to Vp in a quite linear manner, not showing the
pressive strength were linked in a strongly non-linear same ‘plateau’ effect seen with Vp versus c. This is an
manner. encouraging aspect of this ultra-simple test method,
At the extreme low end of the Vp- c spectrum which was adopted for registering the compressive
( c  1 MPa to 10 MPa), the laboratory data for Tertiary strength of (fresh or weathered) rock-joint walls (JCS)
mudstones and sandstones from Japan given in Figure in the shear strength criterion of Barton and Choubey,
2.15 shows roughly an order of magnitude increase in 1977.
28 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(b) (c)

(d) (e)

Figure 2.16 (a to e) Effects of weathering at four sites in Japan cause huge ranges of porosity, strength and P-wave velocity. Saito, 1981.
Environmental effects on velocity 29

(a) (b)

(c)

Figure 2.17 Top: a), b) effect of water saturation on Vp. Bottom: c) Dry Vs/Vp trend over a wide range of Vp. Saito, 1981.

The significant differences of behaviour caused by The Vp/Vs ratios that Saito derived from many hun-
porosity reappear when degree of water saturation and its dreds of data are shown in Figure 2.17c. These particular
effect on Vp are shown side-by-side in Figure 2.17a and data are for dry samples. Vp/Vs ratios are seen to reduce
2.17b. The higher porosities corresponding to higher from about 2.0 at low velocity to about 1.6 at high vel-
weathering grades show very strong (even 200–300%) ocity, broadly following the trends discussed in Chapter 1.
increases in Vp from initial low values, as saturation An example of the effect of saturation is given a sim-
exceeds about 85%. Much less sensitivity to saturation ple theoretical basis by Grainger et al., 1973. Their eval-
(just a weak linear effect) was seen for the fresher, low uation of chalk foundation qualities at the proposed site
porosity, high Vp samples, where Vp increased from just for a proton accelerator facility in Norfolk, England
5 km/s to 5.5 km/s with saturation, in the case of a low revealed one anomalous result, when low quality chalk
porosity sample. (grade V), which was normally sampled above the water
30 Rock quality, seismic velocity, attenuation and anisotropy

table (Vp  0.7 km/s) showed a velocity of 1.95 km/s at structureless melange of angular fragments set in a matrix
one location. The grade V chalk was described by the of deeply-weathered, remoulded chalk, analysed by Grainger
authors as a structureless melange of angular fragments et al., 1973.
set in a matrix of deeply-weathered, remoulded chalk.
A matrix version of the time average equation of
Wyllie et al., 1956, was used by the authors to explain 2.5 Combined effects of moisture
this anomaly as follows: and pressure

1  1  We have seen from Saito’s 1981 data, the importance of


  (2.1) the degree of saturation on Vp in the absence of pres-
V Vfl Vsd
sure (Figure 2.17a,b). In fact, the combined effect of
degree of saturation and stress level have significant influ-
where Vfl  velocity in fluid, ence in rock engineering projects due to the common
Vsd  velocity in solid, and ‘environmental changes’ that are introduced when we
  ratio of the path length in the fluid to excavate a tunnel, slope or foundation, causing changes
the total path length (i.e. the ‘porosity’) of stress (especially unloading) and changes in the pore
The authors assumed Vsd  2.3 km/s for the intact pressures and drainage routes in the so-called ‘excav-
chalk fragments, and first assumed dry conditions with ation disturbed zone’ (EDZ), which will be reviewed in
Vfl  Vair  0.33 km/s, when substituting the meas- detail in Chapter 7.
ured value for grade V  0.7 km/s in equation 2.1, to The excavation process (blasting, boring, ripping,
give   0.29. Returning to the assumed saturated etc.) causes stress redistribution and the release of radial
conditions, using the calculated ‘porosity’ of 0.29 and stresses. Monitoring of Vp in such zones sometimes shows
Vfl  Vwater  1.44 km/s, the authors calculated a areas of increase, but more usually significant reduc-
P-wave velocity for the saturated chalk of 1.97 km/s, close tions in velocity, especially when the rock mass is signifi-
to that measured by the shallow refraction seismic. cantly jointed or damaged by the high tangential stresses,
An interesting variant of the above time average and by the excavation process itself, particularly if by
equation is illustrated for the case of jointed rock, by the less careful versions of drilling-and-blasting.
manner in which joints are assumed to change the seis- It is usually assumed that the release of radial stress
mic velocity. McDowell, 1993, presented the classic and the formation of new fractures by blast gasses are
equation of Wyllie et al., in the form: the two chief causes of velocity (and modulus) reductions
in the EDZ. However, the results of tests on the effect of
L nw L  nw saturation in weathered and micro-cracked materials
 
Vp (rock mass) Vp (joint filler) Vp (rock matrix) (Saito, 1981 and Kelsall et al., 1986) seen in Figures 2.7
(2.2) and 2.17, obviously suggest a third mechanism of vel-
ocity reduction, namely drying out. The presence or
where L is the path length absence of stress and the dry or saturated state create the
n is the number of joints largest ‘environmental’ changes to Vp besides weathering
w is the average width of the joints state (n%, , c).
In practice the velocity through dry jointed rock is Nur and Simmons, 1969, classic experiment with
lower than given by the above, even when the velocity repeated measurements of Vp on a sample of Chelmsford
through air (Vp  330 m/s) is taken into account. This is granite (n  1%) showed a reduction of Vp from
because an air-filled joint will tend to act as an acoustic 5.4 km/s (when the sample was saturated) to a value of
barrier, except round its ends or across points of con- about 3.9 km/s after four days at room temperature, while
tact. The actual travel distance has increased, but (L) drying out. The rapid change in the first 5 hours, even
has not been corrected in the above equation. though the rock is of low porosity, is seen in Figure 2.18.
In the case of water-filled (Vp  1.44 km/s) or clay- Nur and Simmons, 1969, data also show very strong
filled joints, the formula is likely to be more correct, sensitivity to confining stress, especially in the case of
due to the improved coupling across the joint walls. The dry samples, which would mean that an unloaded tun-
saturated condition was successfully modelled by the nel wall ( r  0) in a dried-out rock mass would tend to
matrix version of the time average equation, in the above show significantly lower velocities than if still saturated.
Environmental effects on velocity 31

If velocity reductions appear to exceed what one would crystalline rocks (upper half of Table 2.2). A large-scale
expect in relation to reasonable modulus reductions in parallel would be the effect of ‘environment’ (stress and
an EDZ (from Chapter 7), then drying out seems to be a degree of saturation) on jointed rock which we will see
distinct possibility. The tabulations below, that also in other data sets in later chapters, in particular the data
belong with high stress data from Part II, show the connected with EDZ experiments (Chapter 7).
potential strength of such effects, in comparing dry and To conclude this section on the combined effects of mois-
saturated samples. (Data extracted from Nur and ture and pressure, some data sets will be ‘borrowed’ from
Simmons tabulations, and rounded). future topics in this book, namely the higher pressure
The very fine grain size in the Solenhofen limestone world of Part II, relevant to petroleum reservoirs and
(0.01 mm) compared to the millimetre-size, or several earthquake related tectonophysics.
millimetre-size grains of the other rocks, and its complete An idea of the eventual non-linear nature of Vp-stress
lack of crack porosity is the reason for the almost com- data, is given in some early King, 1966, experiments
plete lack of pressure sensitivity for this rock. Micro-cracks with hydrostatic loading of sandstones, shown in
are presumably the chief cause of the above sensitivities Figure 2.19 with classic psi and ft/sec units. The water-
to pressure and degree of saturation in the case of the saturated and dry states show classic ‘knee’ shapes, and
velocities that begin to converge at high stress, due to
closure of microcracks. The improved coupling with
water, still gives the highest velocity in the saturated
state. The maximum pressures in King’s experiments
were about 35 MPa.
The strong effect of extreme confining pressure, espe-
cially when these pressures go far beyond the uniaxial
strength of the rocks, is typically illustrated by classic
‘knee’ shaped Vp– 3 curves. Figure 2.20 shows a variety
of behaviours from high pressure laboratory test results
on dry samples, given by Wepfer and Christensen, 1991.
A compressible shale (3.0 to 5.7 km/s) and a porous
sandstone (2.2 to 4.0 km/s) show strongest effects of con-
fining pressure, while low porosity sandstone, dolomite
and limestone show only 200 to 300 m/s increases. The
Figure 2.18 Slow air-drying of a saturated sample of granite reduces authors refer to velocity hysteresis; the effect of pressure in
Vp by 1.5 km/s. Nur and Simmons, 1969. closing cracks in the stress range 0–200 MPa (0–2 kb) is

Table 2.2 Confining pressure and dry/saturation effects on the Vp (km/s) of some hard rocks (Nur and Simmons, 1969).

Confining Pressure (MPa)

Vp km/s Rock type Porosity 0 5 10 20 40

dry Casco granite 0.7% 3.3 4.2 5.1 5.7 6.0


saturated 5.3 5.8 6.0 6.1 6.3
dry Westerly granite 0.9% 3.8 4.5 5.0 5.3 5.6
saturated 5.5 5.6 5.7 5.8 5.9
dry Troy granite 0.2% 4.5 5.7 5.9 6.2 6.3
saturated 5.7 6.2 6.2 6.3 6.4
dry Webatuck dolomite 0.7% 5.0 5.9 6.4 6.7 6.9
saturated 6.4 6.6 6.7 6.8 6.9
dry Solenhofen limestone 4.7% 5.6 5.6 5.6 5.6 5.6
saturated 5.6 5.6 5.6 5.7 5.7
dry Bedford limestone 12.3% 2.6 2.8 3.0 3.4 3.8
saturated 4.5 4.6 4.7 4.8 4.8
32 Rock quality, seismic velocity, attenuation and anisotropy

not matched by equal crack opening when unloading


occurs in this region, and velocity hysteresis results.

2.6 Combined effects of moisture


and low temperature

This chapter on ‘environmental’ effects on Vp would not


be complete without reference to the influence of low
temperature and ice formation on Vp. Construction in
permafrost, and monitoring of the ground-freezing
method, for tunnelling though unstable water bearing
areas under environmentally-sensitive areas such as city
streets, could each benefit from seismic monitoring to
determine the progression or regression of the ice front.
Data given by Timur, 1968, show velocity increases
upon freezing that vary from about 20 to 50% for many
saturated porous rocks, as compared to their P-wave
velocities at room temperature. In general, the largest
Figure 2.19 Effect of dry and water-saturated states on Vp-versus- increases are for the most porous rocks. A shale showed
stress, for a sandstone, King, 1966. (Note: an inter- only 8% increase. Dry rock samples are hardly affected
mediate curve for kerogen, lying just below the by cooling below 0°C.
water-saturated curve, has been removed since not The enormously contrasted temperature-velocity
relevant to Part I). graphs for the dry and saturated states, for the first few

Figure 2.20 High pressure effects on Vp (500 to 1000 MPa) for a variety of rock types. Wepfer and Christensen, 1991.
Environmental effects on velocity 33

(a)

(b)

Figure 2.21 Contrasting effects of low temperature on Vp for Berea sandstone in the dry and wet state, with 1  31.3 MPa in each case.
Timur, 1968.

degrees below 0°C, is nicely demonstrated in Figure of jointed rock, one would expect that the smaller,
2.21. This contrast is due to the different rates of freez- finer tips of cracks and joints would freeze first, due to
ing in pore volumes that have different area/volume the more stationary conditions, making ‘ice-wedging’
ratios. Surprisingly perhaps, the author explains that the such an effective mechanism of weathering in moun-
smallest pores actually freeze later due to less favourable tainous terrain, and in more northerly and southerly
area/ volume ratios. In the ‘macro-discontinuity’ world climates.
3 Effects of anisotropy on Vp

In this chapter the ‘simple’ approach to anisotropy caused and that change of properties are related to the behaviour
by micro-cracks or jointing will be taken, considering of the micro-cracks under load. In Chapter 2 it was seen
principally P-wave, azimuthal anisotropy, and anisotropy how micro-cracked and fissured samples were particu-
caused by stress difference. Besides micro-cracks that may larly sensitive to the degree of saturation, since they are
be aligned due to tectonic history or due to existing or seismically much more visible when dry and unloaded,
applied stress anisotropy, there will be fundamental rea- than when saturated and strongly loaded.
sons for velocity anisotropy in foliated, schistose, layered Nur and Simmons, 1969, reported important results of
or inter-bedded rocks with unequal layer stiffness. When stress-induced anisotropy, noting that the largest velocity
jointing and faulting are included, with the special effects change took place in the direction of the applied stress.
of stress anisotropy on these larger scale features, the Prior to loading, isotropic velocity was usually recorded.
potential causes of velocity anisotropy will be numerous. Nur, 1971, showed how the observed velocity anisotropy
Although velocity anisotropy complicates interpretation, caused by stress effects on crack closure could be mod-
at the same time it also provides important information elled, in fair to good agreement with experimental results.
for a rock engineering project, and of course for a frac- Figure 3.1 shows the relative effects of hydrostatic stress
tured petroleum reservoir, if correctly interpreted. It will (0 to 50 MPa), and uniaxial stress (0 to 40 MPa) on the
be seen that the classic alignment of a dominant joint set compressional wave velocity Vp. Here we are also looking
with the maximum horizontal stress direction is often a ahead into levels of stress appropriate to Part II. Under
cause of a double-anisotropy effect. Both the near-surface uniaxial stress, the velocity increase parallel to the stress
and high stress treatment of P-wave anisotropy, as intro- direction is much greater than the velocity increase per-
duced in this chapter, will be supplemented in Chapter pendicular to the stress, due to preferential closure of
14, by studies at considerably greater depth, principally those micro-cracks that are aligned more or less perpen-
in fractured reservoirs. In Chapter 15 the anisotropy dicular to the applied stress. The effect may be enhancing
information found in shear waves will finally be the focus in situ velocity anisotropy effects since Hmax often tends
of attention, as a lot more information is contained in to be parallel or sub-parallel to major jointing, and vel-
waves that polarize in parallel (fast qS1), and (slow qS2) ocity parallel to these joints is highest, independent of
perpendicular directions relative to the discontinuities. the above ‘intact’ rock effects. Similarly to these micro-
These shear wave components show dispersive, frequency cracks, minor joint sets in situ tend to be closed by the
dependent levels of anisotropy, caused, in principle, by major principal stress, giving a further reason for stress-
the dimensions, density and stiffnesses of the fracturing induced velocity anisotropy at larger scale.
and jointing. There are also those who attribute the shear- Holt et al., 1997, suggested that stress dependent vel-
wave anisotropy at depth mostly to micro-cracks. ocity (caused by micro-cracks) seen in cores taken from
great depth may be mainly a result of coring damage
caused by the release of anisotropic stresses. This stress
3.1 An introduction to velocity dependent behaviour is particularly pronounced at low
anisotropy caused by stresses compared to the virgin stress, such as in a triaxial
micro-cracks and jointing test performed below the original stress state. Above the
previous stress state, the sensitivity to stress change was
It is reported that Maurycy Rudzki, the first Professor of less. Holt et al., 1996 and 1997, observed that there was
Geophysics at a university in Cracow, stated his intention little or no stress dependence when no cracks were formed
to do research on the propagation of seismic waves in in the recovery process, nor was there stress dependence
anisotropic media, in 1896. (Helbig and Szaraniec, 2000). when the rock was loaded (or unloaded) near the original
It has also been recognised since early in the 1900s that stress state. In a limited stress regime around the original
compressive stress affects the elastic properties of rock, stress state, the rock behaved as a linear elastic substance.
36 Rock quality, seismic velocity, attenuation and anisotropy

was also utilised. Systematic measurements were made


along different azimuths to check for anisotropic veloci-
ties as a result of anisotropic in situ stresses. The stress-
relieved cores showed azimuth dependent velocity
reductions of as much as 20% in granite and as low as
1% in limestone. In the case of the granite, the max-
imum anisotropy was consistent with the in situ stress
orientation. Significantly, the difference between Vp
(in situ, stressed) and Vp (core, unstressed) was usually less
than 0.5 km/s when the in situ stress difference ( 1– 3)
was limited to about 10 MPa. These velocity differences
tended to be between 0.5 and 1.5 km/s when the stress
difference ( 1– 3) was as large as 20 to 40 MPa in situ.
The dilation and brittle fracturing that occurs when
rock is highly stressed was found by Rummel et al.,
1978, to be a significant source of P-wave anisotropy.
They utilised a biaxial loading arrangement with fast-
reacting servo-control, to study the development of dila-
tion adjacent to the shear failure surfaces developed in
granite. They found that the P-wave velocity increased
continuously in the direction of maximum compression
in the pre-peak region. In the post-failure region the
P-wave velocity decreased almost reversibly with redu-
cing compression.
By comparison, the minimum and intermediate prin-
cipal stress directions suffered a marked reduction of
P-wave velocity (recorded as travel time increases), after
fracturing was initiated. As they pointed out, enhanced
permeability would be a related phenomen of such dila-
tion: this would presumably occur mostly in the 1
direction. The authors mentioned the need to be aware
of the possibility for increasing velocity anisotropy, when
interpreting field seismic data in crustal regions where
large tectonic stresses are assumed to be operating.
Failure processes in intact rock (Berea sandstone) up to
and beyond the brittle ductile transition, with simultan-
eous monitoring of axial and lateral P-wave and S-wave
signatures were reported by Scott et al., 1993. Confining
pressures of 20 to 138 MPa were used. Slight P-wave
anisotropy at the start of each test (due to a weak bedding
Figure 3.1 Velocity anisotropy under stress (I hydrostatic, II meas- fabric) were strongly enhanced at increasing axial strains,
ured parallel to uniaxial, III measured perpendicular to as micro-cracks tended to close perpendicular to 1, and
uniaxial), due to micro-crack closure perpendicular to open parallel to 1. At failure in the brittle regime, the
stress. Nur, 1971. shear fracture formation caused a small break in the
P-wave signal, followed by constant Vp (axial) and Vp
(lateral) velocities as shown in Figure 3.2a.
A comprehensive in situ and laboratory study was During ductile deformation at much higher stresses,
reported by Engelder and Plumb, 1984, using shallow the P-wave anisotropy continued to increase, presum-
boreholes at in situ sites that were free of joints and ably due to a more pervasive micro-crack and grain crush-
above the water table. Dried core from the same holes ing development. By comparison, the shear fractures
Effects of anisotropy on Vp 37

Figure 3.2 a) P- and S-wave anisotropy as a function of confining pressure level for Berea sandstone samples. Note the effect on the veloci-
ties of the onset of dilatancy, and the fracturing event. b) As the ratio of differential stress to ultimate strength rises, the P-wave
anisotropy is seen to increase, but high confinement removes this anisotropy. Scott et al., 1993.
38 Rock quality, seismic velocity, attenuation and anisotropy

developed at lower confining pressure were only sur- micro-cracks. It is a slightly stronger effect when loading-
rounded by a limited zone of micro-cracks. unloading occurs perpendicular to the fabric, as one
Scott et al., show an interesting plot of P-wave velocity might expect.
anisotropy in relation to confining pressure and differ- A extensive collection of laboratory data that show the
ential stress level (Figure 3.2b) that nicely demonstrates clear effect of the measurement direction in relation to
the increasing anisotropy of Vp (axial) and Vp (lateral) the foliation (0°, 45° or 90°) was given by Tsidzi, 1997.
close to failure (at least 30% drop in Vp lateral) and the He used the ultrasonic pulse transmission technique to
reduction of this anisotropy at high stress levels. derive Vp data for intact samples of amphibolite, gneiss,
hornfels, phyllite, schist, slate and quartzite (the latter
only weakly or very weakly foliated). Both dry and unsat-
3.2 Velocity anisotropy caused by urated conditions were tested. The effects of loading were
fabric not reported. Tsidzi, 1997, suggested that ‘strongly’,
‘moderately’ and ‘weakly’ foliated rocks could be expected
Intact specimens of rock that exhibit strongly anisotropic to show velocity anisotropies of 40–20%, 20–6% and
or orthotropic tendencies such as slate, show significant 6–2% respectively. In Table 3.1, some results have been
velocity differences when measured parallel to foliation selected from the much larger set of data given by the
(e.g., 5.2 km/s) and perpendicular to foliation (e.g., author.
4.2 km/s). Duellmann and Heitfeld, 1978, show that this Strongly foliated gneiss from the Nagra project in
anisotropy varies smoothly as the angle of incidence to the Switzerland showed even stronger anisotropy, giving,
foliation is varied from 0° to 90°, as shown for loading in the dry state, a parallel-to-schistocity Vp value of
and unloading cases in Figure 3.3. The minor velocity 4.4 km/s, and only 3.1 km/s perpendicular to this direc-
hysteresis seen on unloading is presumably due to load- tion (Hesler et al., 1996). Figure 3.4 shows that both
deformation hysteresis of the fabric, or of eventual these extremes were achieved at the lowest axial stress of
about 2 MPa, while the application of more than 25 MPa
appeared to largely remove the velocity anisotropy; the
slow perpendicular direction converging to the fast parallel
direction above this stress level.
This convergence is in direct contrast to the micro-
crack related divergent (increasing anisotropy) behav-
iour shown by Nur’s (1971) results, in Figure 3.1. As
noted in Chapter 2, the effect of saturation is to remove
much of the effect of load increase on velocity, and the
same appears to be the case for anisotropy caused by
fabric. Figure 3.4 shows only slight velocity anisotropy
in the case of saturated samples of gneiss, though the
data set is limited.
An important contribution to the understanding of the
three-dimensional anisotropy of dense shales was reported
Figure 3.3 Velocity anisotropy of intact samples of slate due to cleav- by Zinszner et al., 2002. They used ultrasonic techniques
age. Duellmann and Heitfeld, 1978. in the laboratory, to measure the multi-directional P-wave

Table 3.1 A selection of Vp anisotropy data, showing the effect of foliation, schistocity and cleavage, and the dry or saturated state, when
the velocity measurement direction is parallel, at 45°, or perpendicular to the particular planar fabric. Tsidzi, 1997.

Rock Type   0°   45°   90°   0°   45°   90°

Condition Dry Dry Dry Saturated Saturated Saturated


Gneiss (SW) 5102 4211 3956 5918 5237 5081
Phyllite (F) 6010 5130 5090 6050 5417 5307
Schist (SW) 6641 5802 5151 6706 5932 5378
Slate (F) 5913 5074 4893 5745 4722 4283
Effects of anisotropy on Vp 39

velocity across a 666 cm, 18-sided truncated cube The authors also gave the results of velocity measure-
of the Tournemire shale, whose format is illustrated in ments under uniaxial stress levels from 0 to 20 MPa,
Figure 3.5a, using a slate example. Their interpretation which show a remarkable lack of stress sensitivity: the
of qP velocities is shown in the form of a Wulff ’s stere- Vp – stress curves giving almost horizontal straight lines
ogram, in Figure 3.5b. As may be noted, the minimum between the seven different stress levels applied. However,
velocity of approx. 3,200 m/s is recorded perpendicular the directional effect was marked, possibly accentuated by
to the bedding (Z-axis), while the maximum of approx. a tendency for slight shear in ‘diagonal’ directions of load-
4,250 m/s is parallel to the bedding (X, Y, etc.) ing relative to the bedding. The lowest velocities were in
the ZXX, YZY, ZXZ, YZZ and ZYZ (sub-perpendicular
to bedding) directions giving velocities of only 1,700 to
1,800 m/s, while in the XXX, YYX, YYY, XXZ, YYX and
YYZ (sub-parallel to bedding) directions, velocities were
as high as 4,200 to 4,300 m/s.
Under the level of compression applied, and with
presumed careful preservation of the samples, the ZZZ

(a)

(b)

Figure 3.5 Ultrasonic measurements on a truncated cube of dense


shale recovered from the Tournemire experimental tun-
nel, south of Aveyron, in France. a) Truncated cube
Figure 3.4 Effect of schistocity in a strongly anisotropic gneiss, model of slate, showing axes. b) Interpolation of qP
loaded to 40 MPa, parallel or perpendicular to the fabric. velocity measurements for a sample of the shale, from the
Hesler et al., 1996. west gallery. Zinszner et al., 2002.
40 Rock quality, seismic velocity, attenuation and anisotropy

direction gave intermediate velocities in this case, with this jointing and minimum values of 5.1 km/s more
roughly 3,200 to 3,300 m/s. These were similar to in situ or less perpendicular to the jointing (New, 1985). The
P-wave velocities calculated from seismic tomography, velocity rosette shown in Figure 3.7 is a convenient
where in the vertical direction they recorded 3,125 m/s, way of representing the anisotropy, but the possible rea-
with some reduction to 2,950 m/s in a tectonically dis- sons for some of the other features on the rosette, for
turbed area near a sub-vertical fault. example the marked reduction between 30° and 40° (not
exactly perpendicular to the 120–130° joint orientations)
was not given. Perhaps the principal stress had rotated
3.3 Velocity anisotropy caused by some 20° to 30° to 140–150°, giving a low velocity per-
rock joints pendicular to h (minimum), or shear stress effects were
involved.
Masuda, 1964, gave a simple but illustrative example of Noting the complexity of describing jointed rock
the effect of jointing and joint direction on the masses and their physical anisotropy in relation to
anisotropic velocity of blocks of granite at the Kurobe deformability and seismic velocity, Oda et al., 1986,
IV dam site in Japan. Figure 3.6 shows P-wave veloci- developed a crack tensor technique which they compared
ties in the dry and saturated state, for three orthogonal with laboratory tests on artificially jointed samples, and
measurement directions. Velocity anisotropy was signifi- with in situ tests on jointed granite. The artificial samples
cant and sometimes amounted to 20% or even 25% dif- of gypsum plaster were cast with artificial, deformable
ference in velocity. The slowest direction was of course cracks made of deformable greased paper. In Figure
when crossing the joints, the fastest when parallel. The 3.8a, it will be noted that the cracks have either a ran-
loading state of the blocks was not referred to, but judg- dom distribution or an ordered ‘N-S’ distribution. The
ing by the extreme effect of the dry or wet state, possi- squared velocity ratio (V/Vo)2 which is the measured
bly the blocks were under low or zero load when these ultrasonic velocity normalised by that of the intact sam-
velocity measurements were made. ple (Vo), showed corresponding isotropic or anisotropic
A massive granitic site in Cornwall, England, with distributions.
one set of predominant jointing striking ESE–WNW,
(note rotated axis), caused seismic velocities to be quite
anisotropic, with maximum velocities of 5.5 km/s parallel

Figure 3.6 Effects of measurement direction (and saturation) on Figure 3.7 Vp anisotropy at a massive granite site in Cornwall,
Vp values across jointed blocks of rock, at Kurobe IV England. Vp(max) was parallel to the single set of joints.
dam site in Japan. Masuda, 1964. New, 1985.
Effects of anisotropy on Vp 41

(i)

(ii)

Figure 3.8 Velocity anisotropy of gypsum samples with flaws, and of two jointed granite sites in Japan. Oda et al., 1986.
42 Rock quality, seismic velocity, attenuation and anisotropy

Figure 3.9 Azimuthal Vp anisotropy in jointed limestones at a ‘dry’ (left) and saturated site (right). Bamford and Nunn, 1979.

The graphic results of these authors’ in situ tests are They used radial (20° interval) geophone spreads at
shown in Figure 3.8b. Both granite sites were anisotropic a total of four sites in chalk in Lincolnshire, England. The
and the intensity of jointing differed, as clearly shown chalk was not exposed at the seismic measurement loca-
by the magnitude of the squared velocity ratio (Vo for tions, but two quarries in the area were mapped to obtain
granite samples was 4.5 km/s). The authors’ crack-tensor joint orientation data. Strong velocity anisotropy was
calculation showed remarkably good agreement with measured at three of the four sites (Figure 3.10), and max-
the seismic anisotropy measurements. The orientations ima at between 5° and 25° were found to correspond
of the velocity distributions are clearly dominated by the with dominant near-vertical joints which were perpen-
two ‘fast’ velocity directions sub-parallel to the relevant dicular to the axis of a major monocline, which had a
joint set directions. predominant direction of 15°  7°.
As seen in Figure 3.8a, small uniformly distributed ran- The velocity anisotropy of all four sites is compared
dom cracks cause seismic velocities to be isotropically in Figure 3.10a, and a comparison of velocity anisotropy
reduced in relation to an uncracked matrix. Attenuation and resistivity anisotropy for site CFR is given in Figure
is increased isotropically, and Vp/Vs ratios are also 3.10b. For the case of site RGQ, Vmax. and Vmin. were
changed. In contrast, most jointing shows some overall 2.85 and 1.75 km/s, giving a total velocity anisotropy
alignment, and ensures anisotropic seismic response. (Vmax.  Vmin.)/Vmax.  0.38, i.e., approximately 20%
An analysis of the seismic refraction tests at regularly around the mean of 2.25 km/s. Resistivity anisotropy
jointed sites in limestones (Bamford and Nunn, 1979) (Figure 3.10b) was greater than seismic anisotropy for
given by Crampin et al., 1980, indicated that the vel- the case compared, possibly due to the strong contribu-
ocity anisotropy (shown in Figures 3.9a and 3.9b) was tion of a fluid bearing joint set.
also very sensitive to the degree of saturation of the joints. An in-depth investigation of anisotropy caused by per-
The two maxima (at about 40° and 120°) were clearly sistent sub-vertical jointing at a geothermal site in the
related to two sets of joints that intersected at about 80°. USA (Beaver County, Utah) was described by Leary and
Details of similar seismic refraction tests to those Henyey, 1985. The authors analysed in detail why, if
referred to in Figure 3.9 are given by Nunn et al., 1983. a significant number of vertical joints remained open (due
Effects of anisotropy on Vp 43

Figure 3.10 a) Azimuthal Vp anisotropy at jointed limestone sites in Lincolnshire, England. b) A comparison of Vp and resistivity anisotropy
at one of the four sites (see diamond symbols), is also given. Nunn et al., 1983.

to horizontal stress anisotropy), the elastic properties and incidence relative to the crack plane normals, where
hence the seismic velocities would be anisotropic.   1. Vpo is the velocity without cracks.
Compressional waves travelling perpendicular to the At the geothermal site in question, the authors used
joints would obviously be slowed more than those trav- 22 clusters of shots (sources) each located within 1 km of
elling parallel to the joints. However, the authors cau- the wellhead, and used geophones downhole at depths
tioned that minor geologic structure and mineral fabric ranging from 30 m to about 700 m. The source clusters
could also influence the measured velocity anisotropy. were at about 160 m, 280 m and 370 m from the well-
Following earlier work by Garbin and Knopoff, head along six radial lines. The close-in shots were far
Crampin, and McGonigle and Bamford, Leary and enough from the wellhead that casing or tube waves did
Henyey, 1985, gave simplified equations for the effect not obscure first arrivals.
of cracks (or joints) on seismic velocity, and the effect of The results of these tests are shown in Figure 3.11.
the dominant direction of the cracks (or joints). The They demonstrate both azimuthal velocity anisotropy
following two equations are given for the dry and sat- and velocity-depth effects. The two sets of data shown in
urated states. Ignoring higher order terms: the figure, represent average P-wave velocity for seismic
1) For dry cracks: waves originating 370 m from the wellhead, which were
received at two depth ranges in the well (0–300 m,
VP2  VP20 (1  71/21e  8/3e cos 2 shown as triangles and 460–520 m, shown as squares).
 e/21 cos 4 ) (3.1) The 1.5 to 2.0 km/s increase in velocity is surpris-
ingly large for an average depth increase of only about
2) For saturated cracks: 350 m. However, Barton et al., 1994, showed a similar
velocity increase even in the first 50 m at the Gjøvik
cavern site in Norway, due to several MPa increase in
VP2  VP20 (1  8/21e  8/21e cos 4 ) (3.2) stress in rock that had more or less unchanged fre-
quency of jointing and RQD and rock quality Q, in
where:   Nr3/V is the crack (or joint) density of N these first 50 metres. In other words, the increased
cracks of radius r in a volume V, and is the angle of stress acting on the joints (3 to 5 MPa in this case),
44 Rock quality, seismic velocity, attenuation and anisotropy

Figure 3.12 Ultrasonic and longer wave length seismic investiga-


tion of four joint sets in dolomitic limestones, which
had greater frequency (sets I and II), or lesser fre-
quency (sets III and IV). The strengths of the velocity
anisotropy of the different frequencies of jointing are
distinguished in a logical manner. Lykoshin et al., 1971.

equations (3.1 and 3.2). This topic will be treated fully


in Chapter 5.
When jointing intensity is quite different between the
different sets of joints, the use of different seismic wave
lengths may be important in distinguishing the behaviour
in different orientations. Lykoshin et al., 1971, describe
the use of ultrasonic measurement with wave lengths of
0.8 to 0.1 m, and the seismic method with wave lengths
of 8 to 15 m, for distinguishing the velocity anisotropy
caused by the closely spaced joints (S  0.1 to 0.2 m),
from the velocity anisotropy caused by the much wider
spaced joints.
The results of this interesting, and quite early dual-scale
Figure 3.11 Effects of joint set anisotropy on velocities, with velocity anisotropy measurement are shown in Figure
depth effects superimposed, based on an ‘areal well 3.12. Sets I and II correspond to the joints with the clos-
shoot’ or 3-D VSP measurements at a borehole. Vp est spacing, and Sets III and IV which had lower fre-
anisotropy results are shown in the lower diagram, for
quency, were separately logged by this method.
receivers at depth ranges of 0–300 m (lower, average-
A helpful diagram of jointing, ‘broken down’ into its
velocity curves) and 460–520 m (higher, average-
velocity curves). Leary and Henyey, 1985.
component sets, was presented by Olson and Pollard,
1989. This is shown in Figure 3.13. The figure was used
by Schoenberg and Sayers, 1995 to illustrate their build-
ing of a stiffness matrix for a rock mass. This they did
rather than reduced joint frequency, was the reason for by summing the compliance tensor of an unjointed back-
the velocity increase of 2 km/s at a site saturated nearly ground rock and the compliance tensors for each set of
to the surface. The important effect of stress on joint parallel or aligned joints. This they inverted into the
closure and seismic ‘visibility’ is not treated in the above form of a stiffness tensor, which they suggested was more
Effects of anisotropy on Vp 45

and shear stiffnesses, i.e. Kn  Ks, even for the case of


small laboratory samples of rock joints, whether under
low or high normal stress. This is because Kn
and Ks are entirely different physical deformation processes,
involving the normal deformation of a stiff compressed
joint, and the less stiff shearing deformation along the
same joint. The ‘macro-displacement’ stiffness anisotropy
increases further with increased block size. (Barton and
Bandis, 1982).
However, it is at present uncertain to which degree
this fundamental rock mechanics aspect influences the
dynamic rock physics aspect of micro-displacements that
are presumably elastic in nature. There is controversy
on this aspect, and even on whether friction along joints
is a valid attenuation mechanism, as assumed for so long.
We will attempt to resolve some of these questions in
the more comprehensive chapters of Part II.

3.4 Velocity anisotropy caused by


interbedding

The commonly occurring interbedding (alternation) of


sedimentary strata of different stiffness, such as sand-
stone, shale and mudstone; layers which also have dif-
ferent porosity, density and uniaxial strength, causes
anisotropy of all the major mechanical parameters and
also affects all the components of velocity. (In relation
to petroleum reservoirs, the reader is directed to Part II,
Chapter 14, for a fuller discussion of this fundamental
Figure 3.13 a) Joint traces from a 1 m thick bed of limestone, ‘dis- topic). In this section, some observations of the effects
aggregated’ in b) and c) into their two component of inter-bedding on near-surface civil engineering pro-
sets. Olson and Pollard, 1989. jects will be presented.
Fine layering of sedimentary strata means that the
dominant wavelength of a seismic pulse is long com-
useful in the consideration of elastic wave propagation pared to the thickness of individual layers. The medium
through rock masses, than the compliance, which is the will then exhibit effective (and real) anisotropy, with
inverse of stiffness (as commonly used in rock mechan- a vertical symmetry axis in the case of horizontal layer-
ics). Schoenberg and Sayers went on to apply their rock ing. In the presence of hydrocarbons this layered medium
mass stiffness matrix to the interpretation of shear-wave may show substantial attenuation and velocity dispersion.
anisotropy, which is strictly the topic of Chapter 15. The combination of effective anisotropy and attenu-
An aspect that will reduce the predicted anisotropy ation means Qseis anisotropy and anisotropic velocities
when applying the Schoenberg and Sayers stiffness (Carcione, 2000) as we shall see later.
matrix, is the fact that they assumed equal shear and nor- Oberti et al., 1979, reported a very instructive set of
mal compliances, based on seismic imaging of some per- in situ near-surface measurements that involved down-
haps not ideally suited, roughened ‘lucite-sheet’ models. hole sonic logging, cross-hole logging and comparison
(These will also be discussed in Chapter 15). with deformation moduli determined at different depths
Compliance is the inverse of stiffness. In the ‘macro- below plate loading tests. The latter were performed
displacement’ world of rock mechanics, there is a one to parallel and perpendicular to the strata, and could
two order of magnitude difference between the normal therefore be compared with the anisotropic velocities.
46 Rock quality, seismic velocity, attenuation and anisotropy

Figure 3.14 Seismic cross-hole and downhole investigations of marl-sandstone interbedded strata, at a dam site in Italy. Oberti et al., 1979.

Figure 3.15 Correlated anisotropy for the velocity and ‘static’ deformation moduli, as recorded in the interbedded marl-sandstone sequences
shown in the previous figure. Oberti et al., 1979.

The rhythmically layered sandstone and marl, with a anisotropy in this orthotropic rock mass was 4.3 km/s
dip of 27°, formed the foundation for an arch-gravity (perpendicular to the layers) and 5.0 km/s (parallel to the
dam in the Apennines in Italy. layers). Differences can be noted between the higher
Figure 3.14 illustrates the geological sequence and velocities in the sandstones and the lower velocities in
location of boreholes. The exploratory tunnel used for the marl.
the plate loading tests, shown in Figure 3.15, was at Figure 3.15 shows a comparison of the sonic measure-
30 to 35 m depth, and ran parallel to the strike of the ments performed in central boreholes beneath each plate
inter-bedded strata. loading location, where deformations were also recorded
The three boreholes (A1 to A3) were parallel and with extensometers, so that ‘static’ deformation moduli
spaced at 3 m centres. Sonic and cross-hole logs are could be calculated at different depths. The lower static
shown sequentially in Figure 3.14. The mean velocity moduli and lower velocities of the disturbed near-surface
Effects of anisotropy on Vp 47

re-visited in several contexts in Part II, particularly


regarding frequency dependent and stress dependent
attenuation, described in Chapter 10.

3.5 Velocity anisotropy caused by


faults

A final category of anisotropy that will be described in


this chapter is that caused by major faulting. An instruct-
ive example is provided from Japan, where Ikeda et al.,
1981, describe some of the extensive Japanese high-speed
railway (Shinkansen) tunnel investigations in major fault
zones. They show characteristic variations in velocity in
a major 300 m wide faulted zone in the Rokkô Tunnel,
with three zones of velocities as low as 2.2 km/s (Figure
3.17a, b, c).
Extensive investigation galleries enabled the authors
to investigate the effect of the angle between the fault
boundary and the seismic investigation line. When the
angle is very acute, and if the fault is also dipping at a
shallow angle ( in Figure 3.17b), a false high velocity
(V) may be registered, or it becomes impossible to regi-
ster the fault.
The authors also assembled seismic data from 100
rail tunnels, with emphasis on fault zones, and heavily
fractured rock. In Figure 3.17c, Vc is the higher vel-
ocity of the competent rock surrounding the heavily
Figure 3.16 Inter-bedded limestones, shales and sandstones in a fractured zone, which is given a velocity classification A
230 m deep well, Sams, 1995. Resolution of detail in to F in the table shown in this lower figure. The heavily
finely interlayered sequences. fractured zone has the lower velocity V. They observed
that the clay core of fault zones could have velocities as
rock are evident, especially that of the marl in the invert, low as 0.8 to 2.5 km/s.
where moisture content perhaps had increased. It is of interest to note that the two parallel fault
As can be seen in Chapter 5, the deformation moduli zones depicted in Figure 3.17a, which have a minimum
and velocities measured in these tests correlate quite velocity of 2.2 km/s, created many months of delay in
closely with the ‘Vp–Q–M’ model, where the rock tunnelling, as can be judged by the profusion of investi-
quality Q-value, or the velocity, are both seen to correl- gation adits into this regional Otsuki Fault zone.
ate with the ‘static’ deformation modulus (M), provided In another rock type, on the other side of the globe,
that appropriate corrections are made for porosity, rock the same seismic velocity of about 2.0 to 2.5 km/s
strengths and depth. allowed tunnelling progress (with TBM) of up to 1500
A case intermediate between the near-surface and metres per month, the reason being the high porosity of
petroleum reservoir interpretation was presented by an entirely different, easily excavated, and relatively stable
Sams, 1995, for inter-bedded limestone, shale, sandstone rock type, when not heavily jointed. The UK-France
sequences in a 230 m deep (and subsequently add- Channel Tunnel chalk marl had a porosity n  27%,
itional) research well. The combined use of a borehole and an easily cuttable strength of only 4 to 9 MPa.
compensated (BHC) sonic logging tool, a compensated The Q–Vp relationship (Vp  3.5  log10 Q) for hard,
formation density tool, and a Formation Micro Scanner non-porous, near-surface rocks presented in Chapter 1,
(FM) was capable of resolving much of the detail of would normally predict a Q-value as low as 0.003 to 0.01
finely interlayered rock sequences. This research will be (exceptionally poor) for such low velocities: as if the low
48 Rock quality, seismic velocity, attenuation and anisotropy

Figure 3.17 Seismic investigations of fault zones at Japanese rail tunnels. Ikeda et al., 1981. a) Plan view of fault zone crossing the tunnels.
b) False high velocity (V) caused by too acute angle of the seismic investigation line relative to the fault zone. Width of frac-
tured zone has small effect. c) Integrated results from 100 rail tunnels giving a velocity ratio expression for the low fault zone
velocity (V) in relation to the surrounding competent rock (Vc).

velocity represented a fault zone. But in the chalk marl in and a high porosity (i.e. n  27%) will be described in
question, Q-values were in the range 10 to 20 where Chapter 5. They are essential for integrating rock quality
these record tunnelling speeds were achieved. The missing Q and Vp – in softer rocks.
corrections for a low uniaxial strength (i.e. c  5 MPa)
4 Cross-hole velocity and
cross-hole velocity tomography

Cross-hole and between-gallery seismic work has been Mratinje Dam in Yugoslavia, as reported by Kujundzíc,
performed for many years at major dam projects, particu- 1979. This figure shows the separate sites of the deform-
larly at the sites of arch dams, where the deformation ation tests for determining E (the dynamic elastic modu-
moduli of the rock foundation and valley walls are of lus), D (the deformation modulus) and Vp (the local
most concern. Unfortunately, the large number of dams value of Vp at the deformability test site).
constructed from the 1960s to the early 1980s did not Some of the cross-hole, between gallery and borehole-
have the benefit of tomographic imaging, in which not to-gallery velocity measurements are shown in more
just the average velocity between source and receiver, but detail in Figure 4.2 (from Ivanovíc et al., 1970), where
also the approximate location and velocity could be dis- the ‘fans’ of velocity can be readily observed. By relating
played, following inversion of the multi-source-multi- the velocity at each test site to the moduli, the larger
receiver-position data. The use of pairs of boreholes (or scale cross-hole results could be used to extrapolate the
multiple holes), for direct access to a ‘hidden plane’ (or expensive and time-consuming tests to other parts of
planes), on which representative velocities and their loca- the foundation. In Chapter 6 we will see some of the
tion could be calculated has many advantages for assess- inter-relationships that have been developed between
ing the severity of fault zones, the need for treatment Vp, Eintact and D, for comparison with Edynamic which
of the ground, or even in some cases the avoidance of can be derived from Vp, Vs and density, as indicated in
bad ground. In this chapter, a wide variety of cross-hole Chapter 1.
seismic tomographic imaging of the sub-surface will be Possible pitfalls when performing cross-hole seismic
reviewed, from tunnels and caverns, to mining pillars, measurements in low velocity layered media such as
blasting-effect analysis, excavation disturbed zone map- clays, which presumably will also affect cross-hole seismic
ping, and analysis of grouting efficiency. tomography in similar media are illustrated in Figure 4.3.
McCann et al., 1975, indicate that there is an apparent
decrease in the velocity of the high velocity layers with
4.1 Cross-hole seismic for increasing separation of the boreholes. First arrivals at
extrapolation of properties the common depth of 7.4 m showed velocities of 2.18,
1.97, 1.83, 1.81 and 1.78 km/s with borehole separ-
In the initial stages of site description for a civil engineer- ations increasing successively from 2.9 m to 15.1 m. The
ing project, geological mapping of major structures may high frequency direct first arrival received at small bore-
be followed by imaging of these features, using large scale hole separations was replaced by a long dispersed wave-
reflection techniques. As emphasised by Cosma et al., form at the largest separations. Attenuation of the higher
2001, subsequent access in a very limited number of holes frequency, higher velocity part of the wave at increasing
will normally suggest VSP measurements, with sources distance was apparently occurring. The authors used a
at the ground surface. When the construction phase is high voltage ‘sparker’ source in their measurements. The
begun, access via a larger number of boreholes, even those strongly attenuating properties of the clay were presum-
drilled from shafts or tunnel walls will allow a combin- ably the cause of this result.
ation of smaller scale VSP, and direct cross-hole tomog- Various seismic wave characterisation methods were
raphy, giving velocity and location. Later in this chapter we compared at a rock anchor foundation site by Ebisu et al.,
will see some exceptional applications of ‘close-in’ seismic 1992. Figure 4.4 shows P-wave data interpreted from
tomography. seismic refraction, downhole logging, cross-hole and sur-
A classic example of cross-hole and between-gallery face exploration. The discrepancies between the systems
seismic is that shown in Figure 4.1 from the 220 m high should serve as a warning that many factors need to be
50 Rock quality, seismic velocity, attenuation and anisotropy

Figure 4.1 Cross-hole and gallery-to-gallery seismic tests at the Mratinje Dam in Yugoslavia, for extrapolating deformability tests.
Kujundzíc, 1979.
Cross-hole velocity and cross-hole velocity tomography 51

Figure 4.2 Classic example of the use of gallery and cross-hole seismics for extrapolating quality and deformation modulus values at the
Mratinje Dam, Yugoslavia. Ivanovíc et al., 1970.

trend is established. A modulus of 0.3 GPa corresponds


to Vr  0.5 km/s, 1 GPa corresponds to Vr  0.8 km/s.
Usually, when comparing cross-hole and downhole
velocity measurements, the downhole sonic probe is
considered to give a small-scale, and usually higher
velocity than the averaged cross-hole result. However,
the small-scale excavation damage zone (EDZ) that
may also accompany a borehole in incompetent rock,
may presumably be the reason for sometimes measur-
ing a lower velocity at the small scale. In Figure 3.14 in
the chapter on anisotropy, the sonic log of Oberti et al.,
1979, generally showed about 0.5 km/s lower velocity
than the cross-hole result, where the hole spacing was
3 metres. The sonic log also showed greater sensitivity
Figure 4.3 Dispersion effects in layered Oxford clay give apparent
reduced Vp for stiff layer with increased cross-hole sep-
to the layering (marls and sandstone) in this case.
aration. McCann et al., 1975. The large reduction of the velocity in the ‘within-
the-borehole’ measurement shown in Figure 4.4 (Ebisu
evaluated when interpreting the results of a suite of tests. et al., 1992) was not discussed by the authors, but is
Ebisu et al. prefer the use of the surface wave velocity Vr perhaps an expression of damage caused by the drilling/
to correlate with modulus of deformation. They show on flushing process in these near-surface weathered mater-
a log-log plot of modulus versus Vr, that a consistent ials at the rock anchor foundation in Japan.
52 Rock quality, seismic velocity, attenuation and anisotropy

Figure 4.4 Contrasting P-wave velocities at a rock anchor foun-


dation, using four methods of measurement. Ebisu
et al., 1992.

A more ‘normal’ comparison between a downhole sonic


log and a cross-hole log is that shown in Figure 4.5 from
Whiteley, 1990. Hole spacing was 40 m. The latter shows Figure 4.5 Comparison of a sonic log and a cross-hole (mean
a smoothed, average behaviour. While general trends are velocity) log. Whiteley, 1990.
seen to be remarkably similar, details between the two logs
clearly differ due to the change of scale and location.
source of dynamic energy in an adjacent hole. The artifi-
cial source can be a 1 gm detonator cap, a downhole ham-
4.2 Cross-hole seismic tomography mer, a sparker or a piezoelectric high frequency source,
in tunnelling which is moved successively down the sender hole. The
same spacing of sources and receivers (e.g. 2.5 or 5 m), is
The system of seismic data analysis used in tomo- normally used. The inversion of the travel times of the
graphic studies was probably adapted from the medical multiply crossing ray paths, into velocities, or into other
profession, although the use of ‘superficial’ seismic seismic attributes, is organised in principle into a regular
sources (earthquakes) for inversed imaging of the internal grid in which average solutions for the local velocity (or
structure of the earth seems to be a possible source of other seismic attribute), are produced.
inspiration. The efficient data handling and graphic Tomographic plots of velocity, amplitude and velocity
presentation techniques represented by the tomographic difference are commonly employed. Most frequently,
method, were rapidly adopted in rock engineering pro- the method is used from single pairs or multiple pairs of
jects since roughly the mid-1980s, and to an increasing boreholes drilled from the surface or from the face of a
degree in petroleum engineering. tunnel, to image a pending (or already intersected) fault
The simple principle of the method is that a string of zone. Increasingly in recent years the method is also being
receivers (hydrophones or 3D accelerometers) suspended used in mining to delineate highly stressed and burst
(or pushed into) a borehole at (e.g. 2.5 or 5 m) regular prone areas, which seem to be most closely associated
spacing, are used to receive the seismic signals from a with steep gradients of velocity, where high shear stresses
Cross-hole velocity and cross-hole velocity tomography 53

may be present. The seismic tomography method can be and widening of the pilot tunnel had been performed, to
used remotely and safely in hostile environments, to create a cylindrical oil storage cavern. The initial refrac-
image highly stressed regions of a mine or overstressed tion seismic survey from the pilot tunnel had indicated
rock around a deep tunnel. (The interesting use of a range of velocities of approximately 2.3 to 2.6 km/s,
‘passive’ sources such as acoustic emission (AE) will be representing generally uniform conditions. Laboratory
illustrated briefly in Chapter 7, where average velocities samples of the 10 MPa chalk had indicated a mean
can be calculated.) P-wave velocity of 2.4 km/s at natural water content
Figure 4.6 show some potential layouts for the borehole (13 to 14%), and 2.5 km/s when fully saturated.
arrays. A moving source, for example mining equipment, An example of the potential benefits obtained from
can also be used to obtain a tomographic image, if a suit- cross-hole seismic tomography at a near-surface cavern
able array of receivers is in place, and if measurements are site is shown in Figure 4.7a and b. The measurements
repeated at regular intervals over a suitable length of time. were performed for the Gjøvik Olympic cavern site
Westman et al., 1996, utilised a long wall shearer in an investigation in Norway in 1990. The position of the
Appalachian coal mine in the USA, and sampled this planned, 62 m span, 140,000 m3 cavern was adjusted in
source at 1⁄2 to 1 minute intervals during mining shifts, order to penetrate as little as possible of the lower
while the shearer was moving. Their receivers were geo- velocity, near-surface zone (Vp  4.1 to 4.3 km/s).
phones fixed to rock bolts in the mine entry roofs, close to This was proved in later cavern logging to have rock
the mining face. They produced attenuation tomograms quality Q-values as low as 2 to 5 at the shallow end of the
that changed with time as mining advanced in response to cavern. This quality results from a low to moderate RQD
high stress anomalies, stress release phenomena, changed (frequent smaller pieces of core 10 cm long), up to four
degrees of jointing and stress induced fracturing. joint sets (Jn  12–15), and with some alteration of the
The assumption is often made that P-waves have trav- joint walls (Ja  2). Positive aspects were considerable
elled directly from source to receiver, and a straight line joint roughness (Jr  2 to 3), and surprisingly high hori-
tomography program is used. Curved ray path tomog- zontal stresses. (See Appendix A for Q-parameter ratings.)
raphy is preferred to allow for velocity anisotropy and for These moderate velocities fit the hard rock, near-surface
refraction (McDowell, 1993). relation Vp  3.5  log10 Q presented in Chapter 1 quite
By, 1987, described a comprehensive layout of vertical closely, for the relevant shallow conditions (approx.
boreholes for cross-hole seismic tomography, which was 25 m depth). At the other, deeper end of the cavern, Q-
performed in Oslo for a difficult, faulted section of the values also fell to 2 or 3. Significantly, this rock quality
twin tube, 13 m span Fjellinjen road tunnels (Figure 4.6a). Q was lower than the Vp values would have indicated,
Some 20 m of soft clay underlying downtown Oslo had to with this shallow seismic relationship. The fundamen-
be protected from groundwater pressure drawdown. At tal need for depth or stress adjustments in a Q–Vp–M
one location, the rock cover over the arch consisted of only (static modulus of deformation) relationship, are dis-
3 to 5 m of crushed alum shale (damaging to concrete), cussed in Chapter 5.
beneath 20 to 30 m of soft clay, in a major regional fault The details of NGI’s cross-hole tomography, analysed
zone. Selection of freezing for one of the tunnel tubes was in more detail in Chapter 5, indicate a continuous rise in
made on the basis of the seismic results, which were based velocity down the 60 m deep boreholes (approximately
on cross-hole measurements from five boreholes of 60 m from 3.5 km/s to 5.5 km/s), despite more or less constant
depth and a total of eight cross-hole sections. joint frequency, RQD and rock quality Q-values down
In contrast to this layout of vertical holes, Hope et al., the lengths of the recovered rock cores.
1996, working in chalk, used single holes drilled radially This is a good example of stress effects on in situ Vp val-
into the wall of a pilot tunnel, and the upper and lower ues, since hydraulic fracturing stress measurements had
walls of a tunnel, to give two triangular shaped spreads shown h min (and the elastic theory estimate of H max),
(see Figure 4.6b). They obtained a distribution of veloci- to be about 3 and 5 MPa respectively, at cavern depth, i.e.
ties ranging from 1.8 to 2.5 km/s between the crown equivalent to depths of 100–200 metres, if vertical
positions (2 m intervals) and the borehole, and 1.9 to stress alone had been responsible for the rise in Vp.
2.3 km/s between the invert positions (2 m intervals) and Shifting to another category of seismic tomography
the borehole. Lower velocity zones were consistent with applications for tunnelling, it is interesting to note
additional jointing associated with a listric fault cutting that deviated boreholes are quite frequently used in
through the chalk. This was verified after benching down combination with sea-bottom hydrophones to obtain
54 Rock quality, seismic velocity, attenuation and anisotropy

(a) (b)

Figure 4.6 a) Cross-hole tomography arrays to characterise a fault zone at the Oslo Tunnel. By, 1987. b) Some examples of seismic arrays, and
a triangular Vp tomogram for chalk at a storage cavern site in Israel. Hope et al., 1996.
Cross-hole velocity and cross-hole velocity tomography 55

Figure 4.7 Cross-hole velocity tomography performed by NGI for the Gjøvik cavern site investigation, Norway. Note the use of different
velocity scales, to improve the velocity information a) above the cavern, b) at cavern depth. Barton et al., 1994.
56 Rock quality, seismic velocity, attenuation and anisotropy

Figure 4.8 Sub-fjord borehole to seabed seismic tomography, using differently inclined and deviated 250 to 300 m holes drilled from the
land and from an intervening island, and seabed geophones. (Westerdahl and Cosma, priv. comm.1998.)

information about major faults known to be sub-paral- seismic hard to use, as long geophone arrays may receive
lel to these deep fjord depressions. There are numerous shortest path direct waves earlier than the refracted head
sub-fjord and sub-sea tunnel sites in Norway, that have waves. There is also gradational, progressive weathering,
utilised seismic tomography for the near-land part of rather than distinct layering, with less clear development
their routes, with sub-sea refraction measurements for of head waves.
the less accessible kilometres of these tunnels. Consequently, the authors report widespread use of
A typical case is shown in Figure 4.8, for planning tomographic inversion techniques for their tunnels
the 1997–1999 tunnelling under the Oslo Fjord, near through steep terrain, and use not only of hole-to-hole,
Drøbak in southern Norway (pers. comm. Westerdahl, but also hole-to-surface and surface-to-surface configur-
NGI and Cosma, Vibrometric, 1998). In this particu- ations of sources and receiver strings. The authors also
lar case the 50 to 70 m thick, fjord-bottom sediments, show the parallel use of downhole logging, with Vp and
caused some reduction in resolution. The fault zones Vs based interpretation of the three elastic moduli and
were correctly predicted and later encountered in the velocity-depth gradients. The additional use of rock qual-
tunnel, but some details of their structure could not be ity RMR and Q-value core logging, and the extrapolation
detected as well as expected. and intrapolation of properties afforded by the geophysics,
Because of uncertainties about the likely quality of provides a quite reliable basis for tunnel support strat-
difficult pre-grouting into a boulder and clay filled egies, with key attention to fault zones and portal areas.
depression in the bedrock, against a back-pressure from Other near-surface uses of cross-hole seismic tomog-
more than 100 m of seawater, a deeper by-pass tunnel raphy that can be mentioned in this section are of course
was excavated to maintain schedule. Penetration of the the possibilities of using geotomography at dam sites. It
major zone was prepared with more extensive (but is easy to imagine the benefits of correctly located low and
partly unsuccessful) pre-grouting, followed by a time- high velocity zones in a dam foundation such as that
consuming freezing, when the full scope of the situ- illustrated in Figures 4.1 and 4.2, where at that time, (in
ation was understood. The quality of pre-grouting (or the 1960s and 1970s), only average velocities between
that of the freezing process) could probably have been holes could be determined to extrapolate deformation
detected by means of seismic tomography monitoring. moduli across the foundations.
Excellent examples of the use of cross-hole seismic At dam sites that are located in limestones and mar-
tomography, (‘geotomography’) in mountain tunnelling, bles, solution cavities can prove extremely difficult to
are given by Chang and Lee, 2001, who refer to several find and treat with conventional drilling and injection.
tunnels in South Korea, surveyed by these and other Deep, sediment-filled scour-holes in dam site canyons,
methods. The authors point out that severe topographic extending foundation depths many tens of metres could
changes and gradational weathering in mountainous ter- also be mapped more successfully with cross-hole seismic
rain, make the use of conventional travel-time refraction tomography.
Cross-hole velocity and cross-hole velocity tomography 57

By et al., 1988, described the use of the technique at a


dam site in northern Norway. Small, concentrated low
velocity zones of about 3 km/s were observed at numer-
ous locations in the marble-mica schist dam founda-
tions, using cross-hole measurements between seven
inclined boreholes. The information formed a fence-like
picture across the foundation, enabling decisions to be
made about modified grout-curtain designs.
Cross-hole seismic tomography from boreholes drilled
from the surface can also be performed at greater depth,
in order to extrapolate core-logging data to (presently)
inaccessible locations, as a means of optimising layouts
for mines or nuclear waste repositories, or research-
related underground laboratories.
An early example of cross-hole seismic tomography
at the Underground Research Laboratory (URL) site in
Manitoba, Canada is given by Wong et al., 1983. They
first conducted a form of ‘cross-hole VSP’ by fixing a
transmitter in one hole at 100 m depth, while the
receiver was moved by 2.25 m intervals in a vertical
hole 16 m away. In this case both P and S arrivals were
inverted, giving average values of Vp and Vs of 5.5 and
3.1 km/s, and a deduced dynamic Young’s modulus of
65 GPa and a dynamic Poisson’s ratio of 0.245.
Due to the relatively coarse resolution achieved when Figure 4.9 Residual Vp tomogram between two boreholes at the
they subsequently conducted cross-hole seismic tomogra- Underground Research Laboratory (URL), Canada.
phy with a 175 m borehole spacing, they felt the need to These were spaced 175 m apart. The calculation of residual
plot the so-called residual velocity, to accentuate informa- velocity was designed to remove background velocity
tion levels at the location of one of the now quite well (5.5 km/s) and a measured velocity-depth gradient of
2 s1 seen from equation 4.1. Wong et al., 1983.
known, and much researched fracture zones. This inter-
sected borehole URL-6 at 275 m, and was proved to be
the same zone at 300 m depth in borehole M2a, due to
the consistently low velocity between these two locations. years of investigations of the site, cross-hole seismic tomog-
The calculation of residual velocity shown in Figure 4.9 raphy was performed between some of the 1000 to 1200
was designed to remove background velocity and velocity- metre deep boreholes. The layout of the holes and the
depth trends. The authors used piezoelectric transducers results of two of the tomograms are shown in Figure 4.10.
as high frequency sources, in order to improve resolution. Stress levels at 1200 metres depth were as high as
The authors expressed the residual velocity as: 40–50 MPa in the direction BH2 to BH5, and about
20 to 30 MPa in the direction BH2 to BH4. Because of
Vp  Vp  5.5  0.002 (z  100) (4.1)
the fault structures sub-parallel to H (max) seen crossing
with depth expressed in metres (for z  100 m). They the BH2 to BH4 tomogram, and in view of the lower
were thus reducing the measured velocity by 0.2 km/s at horizontal stress acting in this direction, one would
200 m depth, by 0.4 km/s at 300 m depth, making an expect lower average velocities in this tomogram than
assumed velocity – depth gradient of 2 s1. A lot of dis- between BH2 and BH5. This proves not to be the case,
cussion concerning such gradients will be found in and one must therefore question whether the between-
Chapter 11, relevant to continental velocity-depth struc- borehole distance of 600 m is giving similar attenuation
tures, and those derived from mid-ocean, spreading ridge problems as shown in Figure 4.3 earlier in this chapter.
investigations of new basalt crust. The other possible conclusion to draw from the velocity
At the UK Nirex Ltd Sellafield site in NW England, tomograms is that perhaps a deep zone with different
where it had been proposed to locate a low and inter- joint orientation caused the lower velocity region at 800
mediate level nuclear waste repository following many to 1100 metres depth. Whatever the explanation, the
58 Rock quality, seismic velocity, attenuation and anisotropy

Figure 4.10 Deep (1000–1200 m) cross-hole tomography at the UK Nirex Ltd Sellafield site. (Schlumberger GeoQuest, Nirex Report
S/94/007, by kind permission.) (See Color Plate 1)

frequently occurring velocities of about 5.2 to 6.2 km/s 4.3 Cross-hole tomography in mining
relate quite poorly with the mean weighted Q-value of
about 3 (range 1 to 10) which were logged by the NGI Phenomenological results of stress change causing velocity
team of engineering geologists. (Barton et al., 1992a). change will be presented in this section, prior to the in
This discrepancy is most likely due to stress effects on depth review of stress effects on velocities in jointed media
Vp in the jointed (ignimbrite and welded tuff ) rock to be given in Chapter 5. Although the cases reviewed
mass (Barton, 1995). This will be discussed further in are from mining, it may be useful to start with an inter-
the next chapter, where effects of depth and stress in esting high pressure tomography experiment from the
jointed media are reviewed in more detail. Correlations laboratory, described by Scott et al., 1994.
are finally developed between depth and velocity for a The effects of their high pressure loading of an intact
given rock quality Q, also incorporating uniaxial com- cylinder of Berea sandstone with a steel indentor, was
pressive strength and matrix porosity. monitored by 20 acoustic sensors arranged in a ring
Cross-hole velocity and cross-hole velocity tomography 59

Figure 4.11 Ultrasonic tomography to monitor the loading on an indentor on Berea sandstone. a) Experimental set-up, showing acoustic
sensors, load application, and data acquisition. b) Cross-section, showing tomographic plane. c) Acoustic tomograms for dif-
ferent indentor stress levels a to h. Scott et al., 1994.

around the sample. The sample had a porosity of 18%. the indentor at 20.6 MPa applied stress, and finally to
One hundred and thirty ray-paths were analysed to 3.55 km/s at 110 MPa applied stress. The rate of velocity
calculate the velocity in 97 individual elements. The increase declined at higher stresses, presumably due to
experimental set-up is shown in Figure 4.11a,b and the the already reduced pore space. However, the sample
tomographic images for eight load increments (includ- appeared to have remained nearly in the elastic state, and
ing final unloading) are shown in Figure 4.11c. the velocity after final unloading was very similar to that
It was found that the mean velocity of 2.3 km/s for the before loading, except for some increase in the area show-
sample increased to 3.0 km/s a short distance under ing the lowest velocity. Slight damage was assumed.
60 Rock quality, seismic velocity, attenuation and anisotropy

Figure 4.12 Cross-pillar seismic tomography showing ray paths


and Vp values (km/s) across pillars in the Masua Mine
in Italy. The host rock was dolomitic limestone, and
the orebody was mineralized limestone. Barla, 1993.

A second experiment involving a vertical plane of


measurement, revealed velocity increases as before, but
the levels achieved differed from those in the horizontal Figure 4.13 Cross-pillar seismic tomography across a coal pillar,
plane (they were lower) indicating differential stress showing the relation between Vp and the perceived
induced anisotropy. stress level in ‘yield pillar A’. Friedel et al., 1996a.
Following this laboratory demonstration of pressure
effects on Vp in intact (but porous) rock, we can exam-
ine some cases from mining where monitoring of stress Figures 4.13 and 4.14 show the tomographic test
changes was carried out at much larger scale. set-up in each case, and below this the velocities
Barla, 1993, describes the use of seismic tomography (2.25 to 3.75 km/s) and velocity changes (1.5
across three pillars in the ore-body of the Masua Mine in to 2.5 km/s) as a result of the adjacent longwall panel
Italy. While there was a general tendency of high velocity advance. The one day of advance (some 8 metres) caused
(up to 7 km/s) in central parts of the pillars, and lower reductions in velocity, presumably just as required for a
velocities (3–4 km/s) on the outsides of the pillars, there yield pillar function. The local reductions in velocity
was however some variation, and in one perhaps highly probably reflect the adverse effect of an increase in the
mineralised zone, the velocities were highest at one edge vertical pillar stress (‘ 1’), which would cause loosening
of a pillar, as shown in Figure 4.12. and reduced velocity in horizontal directions (as moni-
Friedel et al., 1995, 1996a and 1996b, used seismic tored), where the rock was not well confined.
velocity tomography both in a coal mine (Foidel Creek, Gas and coal outbursts in Polish mines in the Lower
Colorado), and in the deep Homestake gold mine, in Silesian coal basin, and the difficulty or impossibility of
South Dakota, USA for monitoring of apparent stress obtaining test samples due to the fineness of the dis-
changes and stress gradients as a result of mining. At the continuities, led Poldolski et al., 1990, to use time-
coal mine reported by Friedel et al., 1996a, they moni- lapsed tomographic imaging to monitor velocities and
tored velocity and velocity changes in two yield pillars related areas of high stress. The authors describe a 70 ton
alongside active longwall panels. roof fall (and 2600 m3 ejection of methane) and how
Cross-hole velocity and cross-hole velocity tomography 61

McConnell orebody, near Sudbury, Ontario. This is a


steeply dipping sulphide, crossed by a number of devi-
ated boreholes, as illustrated in Figure 4.15a. The author
described the use of a non-destructive piezoceramic vibra-
tor source which was successively lowered down each of
the inclined water-filled holes, with a hydrophone string
of detectors in the nearest neighbouring hole. Figures
4.15b and 4.15c show a schematic of the equipment,
and 1/5 th of the ray-paths between two of the adjacent
holes. The (approximately reproduced) tomogram shown
in Figure 4.15d indicated a clearly delineated orebody
velocity of about 4.0 to 4.5 km/s, compared to the 5.9
to 6.5 km/s of the host rock.
The so-called pixel dimensions for the tomographic
imaging and interpretation were only 1.5 m  2.5 m.
Dominant frequencies were in the 3–4 kHz range.
As a first approximation, a straight ray-path assump-
tion was made to speed the interpretation. As the
author pointed out, actual raypaths were likely to curve
due to refraction in a non-uniform geologic medium,
and could be degraded by false features or artefacts.
Checking of the tomogram structure, using independ-
ent means, including the recovered core, was therefore
advised.
Figure 4.14 Vp as a monitor of increasing stress in ‘yield pillar B’,
adjacent to a longwall-mining advance in coal. Friedel
4.4 Using tomography to monitor
et al., 1996a.
blasting effects

high seismic velocities in the same area correlated with Several investigators have used seismic velocity tom-
increased volumes of coal ejection and degassing from ography to follow the effects of loosening and void
blast holes drilled for shooting the longwall face. The formation caused by blasting. Cumerlato et al., 1988,
link between high stress and high velocity – prior to performed seismic tomographic analysis of pre-blast and
failure, is clear. post-blast quarrying effects in dolomite, in a lime quarry
Friedel et al., 1995, reported monitoring between two in the USA, using a modified refraction seismic tech-
levels of the deep Homestake gold mine in the USA. nique. Figure 4.16 shows pre-blast and post-blast velocity
Their results indicated a sensible correspondence between distributions, and clear advantages of a modified blast
low velocity zones and back-filled areas, ore chutes, and hole loading factor for controlling fracturing. High
so on. High velocity gradients were interpreted as loca- velocity zones (Vp  4.5–6.0 km/s) were reduced to low
tions of potential rock burst. We shall see more examples velocity (0 to 3.5 km/s) when blasting performance was
of the effects of high stress on velocities, when reviewing unfavourable, due to all the crushing and void formation.
the work that has been done in excavation disturbed Maxwell and Young, 1993, used a velocity difference
zones (e.g. Cosma et al., 2001) in Chapter 7, and also image technique for analysing the effect of an explosive
see the possibilities of using acoustic emission (AE) as a detonation in a borehole in granite. The experimental
remote method of monitoring high stress gradient prob- set-up is shown in Figure 4.17a and b. The velocity dif-
lem areas. ference images, examples of which are shown in Figure
Cross-hole seismic tomography has also been in use 4.17c and d, are computed from before-and-after-blasting
to delineate the detailed structure of orebodies, beyond time-delays, along common ray paths. The authors
what can be achieved by intermittent core drilling. A observed extension of the lower velocity zone away
good example was described by Wong, 2000, from the from the blast hole, sub-parallel to the trace of assumed
62 Rock quality, seismic velocity, attenuation and anisotropy

(a) (b)

(c) (d)

Figure 4.15 a) Sulphide orebody delineated by boreholes. b) Schematic of equipment and acquisition geometry (shown vertical). c) One-
fifth of the total 4,200 raypaths for one pair of holes. d) Approximate velocity tomogram, showing the lower velocity of the sul-
phide orebody (Vp  4.0–4.5 km/s), compared to the host rock (Vp  5.9–6.5 km/s). Wong, 2000.
Cross-hole velocity and cross-hole velocity tomography 63

Figure 4.16 Pre-blast (left) and post blast velocity tomograms. Cumerlato et al., 1988.

(a) (b)

(c) (d)

Figure 4.17 a,b) Cross-hole tomography set-up, for monitoring blasting effects in a borehole in granite. c,d) Velocity difference tomograms
showing reduced velocity caused by blasting. Error tomogram on right. Maxwell and Young, 1993.
64 Rock quality, seismic velocity, attenuation and anisotropy

joint planes. Secondary changes may have been associated blast induced fracturing had occurred. These low velocity
with changes in the water table. zones (for example Figure 4.19) showed increases in
Seismic tomography for controlling blast fragmenta- velocity when fluid was injected. These were also the
tion results for mine areas where in situ leaching was areas where lost circulation occurred when drilling
planned, were described by Thill et al., 1992. The results was performed. The joint aperture changes and new
shown in Figures 4.18 and 4.19 were obtained from cross- fractures presumably created poor acoustic coupling so
hole measurements by the US Bureau of Mines at the were readily detected as velocity reductions, later to be
experimental Edgar Mine in Idaho Springs, Colorado. partly recovered when there was resaturation in the area.
They found good correlation between pre-blast and
post-blast velocities that corresponded to areas where
4.5 Alternative tomograms

Thill et al., 1992, show a method they developed of con-


straining the seismic parameters (e.g., velocity) at a com-
mon borehole axis, when two non-coplanar tomograms
were to be joined in a ‘fence’ type presentation, as illus-
trated in Figure 4.18. One result that is not immediately
obvious is the Vp/Vs ratio tomogram shown in Figure
4.19. Why the post-blast high velocity (undamaged)
areas should have the highest Vp/Vs ratios (i.e. 1.88 or
higher) while fractured zones with low velocities have the
lowest ratios, must presumably be due to the changed
saturation, since if each area were saturated the opposite
Figure 4.18 Pre-blast ‘fence’ tomogram at the USBM Edgar Mine result would be expected, as we saw in Chapter 1.
in Colorado, USA. Thill et al., 1992.

Figure 4.19 Post-blast tomograms at the stope leaching site, Edgar Mine, Colorado, USA. Thill et al., 1992.
Cross-hole velocity and cross-hole velocity tomography 65

Another factor also seen in Chapter 1, is the basic, mining site (the Kamioka Mine in Japan). The geolog-
and theoretically determined correspondence between ical setting and source and receiver locations are shown
high Vp/Vs ratios and high dynamic Poisson’s ratios. in Figure 4.20a and b. The P-wave velocity tomogram
The general form of both these tomograms is seen from shown in Figure 4.20c indicates high velocities, even in
Figure 4.19 to be similar, following the theoretical basis locations where ‘fractures’, ‘basic dike’ and ‘fault’ are
given in Chapter 1. shown, suggesting high stresses and reduced sensitivity
In a nice example of the capabilities of alternative to jointing and faulting. The authors therefore utilised
tomographic descriptions of a site, using different seis- amplitude attenuation tomography (Figure 4.20d) and
mic wave form analyses, Watanabe and Sassa, 1996, pulse broadening tomography (Figure 4.20e), which
give three tomographic plots of the same experimental correlated better with the geologic structures.

(c)

(a)

(d)

(b)

(e)

Figure 4.20 a) to e). Three tomograms comparing P-wave velocity, amplitude attenuation and pulse broadening methods of analysis at the
Kamioka Mine in Japan. Watanabe and Sassa, 1996.
66 Rock quality, seismic velocity, attenuation and anisotropy

Amplitude attenuation was calculated for the first rise time or pulse width of the first arrival P-wave
arrival P-wave, and was based on the fact that amplitude is used.
decreases by geometrical spreading, and is a function of
the distance between the source and receiver. The attenu-
ation coefficient (␣) is given by: 4.6 Cross-hole or cross-well
reflection measurement and
f time-lapse tomography
 (4.2)
QV
Although strictly outside the scope of Part I of this book,
which deals mostly with civil and mining engineering
where (Q) is the seismic quality factor, (f ) is the fre- topics, an exception will be made here, concerning
quency and (V) is the velocity. a description of the obvious benefits of using cross-well
Watanabe and Sassa, 1996, suggested that the seis- seismology in the petroleum industry. Paulsson et al.,
mic Q-value was an inherent parameter of the medium 1993, recommended not only cross-hole tomography
that was independent of frequency in the seismic wave but also well-to-well reflection measurements, as shown
frequency range. At the same mine site they listed the diagrammatically in Figure 4.21, to obtain a better
following Vp and Q-seismic values for cores. understanding of the (increasing-as-time-goes-by ?) het-
erogeneity of reservoirs.
Table 4.1 Seismic velocity and Q-seismic of rock cores measured They demonstrated how repeated (time-lapsed) sur-
in the laboratory (Watanabe and Sassa, 1996). veys could be used to follow the progress of enhanced oil
recovery (EOR) programmes, such as steam injection,
Velocity (m/s) Q-seismic
and also pointed out the advantages of the downhole
Gneiss 5700 79 location of both source and receivers, since the attenuat-
Limestone 5470 22 ing weathered (or soft-sediment) layer is no longer limit-
Skarn 4900 28 ing the high frequencies that can be recorded.
Basic dyke 5170 36

The gneiss showed the highest Vp (5.7 km/s) and the


highest Q-seismic value (79). In situ, the gneiss
between points 5 and 20 (Figure 4.20b) had an average
velocity of 5.53 km/s and a Q-seismic value of 16.
The amplitude attenuation tomography shown in
Figure 4.20d gives the Q-seismic values. A dark colour
corresponds to high Q-seismic values of low attenu-
ation. Light colours correspond to the low Q-seismic
values associated with the fault. Soft limestone and
areas oozing water reportedly also showed lower Q-seis-
mic values (i.e., about 10 or less). There seems therefore
to be more evidence here of a fairly close implicit rela-
tionship between the Q-value (the rock mass quality
rating of Barton et al., 1974) and the Q-seismic value.
This will be explored in greater detail in Part II,
Chapter 10.
The final tomographic plot shown in Figure 4.20e is
called pulse-broadening tomography. The dark colour
denotes a large broadening factor, or low attenuation.
The pulse broadening technique is based on the fact
that the wavelength lengthens and the frequency Figure 4.21 Cross-well transmission and reflection tomography for
reduces as a seismic wave travels in the rock mass. The petroleum reservoir definition. Paulsson et al., 1993.
Cross-hole velocity and cross-hole velocity tomography 67

Their results showed the strong correlation between oil


saturation and velocity (e.g., 1.5 km/s at 20% oil satur-
ation to 2.7 km/s at 60% oil saturation). They also noted
the high velocity zone that developed when injecting cold
water, due to resaturation, a result that could presumably
locally reverse in a jointed reservoir, if effective stress
reductions and joint aperture increases exceeded the other-
wise positive effect of resaturation on Vp.
In a keynote lecture at the 6th IAEG Congress in
Amsterdam, Whiteley, 1990, gave particular emphasis to
the high resolution, cross-hole reflection imaging tech-
nique. The three diagrams given in Figure 4.22a, b, and c
illustrate the basic field set-up and two of the methods
(yo-yo and beam steering) for imaging targets from mul-
tiple positions.
In this particular application from Australia, interest
was focussed on the location of shallow coal seams and
of unfavourable structural features that would affect
mining operations. A modified marine sparker source
was lowered in one hole, and an array of closely spaced
marine hydrophones were located in an adjacent hole,
which could be up to 150 metres away. Borehole depths
were up to 300 metres.

Figure 4.22 High resolution, cross-hole reflection imaging technique,


showing the yo-yo and beam steering methods, for delin-
eating shallow coal seams in Australia. Whiteley, 1990.
5 Relationships between rock
quality, depth and seismic
velocity

Efforts to relate rock quality and seismic velocity have the field are being compared with the higher frequency,
been made at intervals, during the development and typically ultrasonic measurements of the laboratory.)
integration of rock engineering and engineering geology.
Included in this review will be RQD, joint or fracture  V 2
frequency (Fm1), and the Q-value (the ‘static’ rock mass RQD%  100  field  (5.1)
 Vlab 
quality rating). Their various relationships to P-wave
velocities obtained from shallow refraction seismic, and
Table 5.1 Relationship between rock quality, RQD and velocity
also from down-hole sonic logging, will be explored.
index, Deere et al., 1967. (VF  field value of Vp,
The correlations obtained have had emphasis on hard VL  laboratory value of Vp).
rocks, with or without weathering, without the compli-
cation of matrix porosity variations, or large ranges of Quality description RQD (%) Velocity index (VF/VL)2
strength and density. These preliminary empirical cor- Very poor Less than 25 0–0.25
relations between RQD and velocity ratio, and between Poor 25–50 0.25–0.5
Fm1 or the Q-value and Vp, must necessarily include Fair 50–75 0.5–0.75
the effect of depth or stress level on Vp, for them to be Good 75–90 0.75–0.9
more widely applied. On the basis of numerous reviews Excellent Over 90 Over 0.9
of deeper seismic measurements, a method is developed
in this chapter, that includes matrix porosity and rock
Other authors, reviewed by McDowell, 1993, have
strength besides all the rock mass attributes of jointing,
suggested the following evaluation of rock quality, as
faulting, weathering and clay. To these are added the
expressed by RQD.
all important influence of depth or stress level, causing
gradual or rapid closure of many or all of the joint sets. Table 5.2 Seismic evaluation of Rock Mass Quality (see McDowell,
1993). The ratios are field-seismic/lab-ultrasonic.

5.1 Some preliminary relationships Joint VF  V 2


 F 
between RQD, F, and Vp Quality frequency
VL  V 
description RQD (%) (m1) L

Relationships between Vp (lab, therefore intact) and Vp Very poor 0–25 18 0–0.4 0–0.2
(field, therefore jointed) have been suggested as a seismic Poor 25–50 15–18 0.4–0.6 0.2–0.4
measure of degree of jointing for many decades. Deere Fair 50–75 8–5 0.6–0.8 0.4–0.6
et al., 1967, found that the ratio Vfield/Vlab. when squared, Good 75–90 5–1 0.8–0.9 0.6–0.8
was numerically very close to the value of RQD (expressed Excellent 90–100 1 0.9–1.0 0.8–1.0
as a ratio rather than a percentage), at least for near-
surface measurements. (RQD is defined as the % of The above sets of relationships are only approximate,
core that has core sticks 10 cm long, for selected as too few factors that obviously affect Vp values for the
structural domains, or for specific lengths of core). The rock mass are actually ‘captured’ in the RQD value
following simple table shows the central trend of this alone. RQD on its own is an insufficient descriptor of
relationship, which however shows considerable scatter. the rock mass quality. However, as a single parameter
(It should be noted that seismic refraction velocities in it is very effective in heavily jointed rock masses, where
70 Rock quality, seismic velocity, attenuation and anisotropy

it is particularly sensitive to the state of weathering or


alteration, since the frequent occurrence of clay-fillings
in the accentuated near-surface jointing, gives appro-
priately low values of RQD, for example 10 to 20%.
(Note that ‘incompetent’ rock pieces that can be broken
by hand are excluded, i.e. local RQD  0%, following
Deere’s recommendations.)
When joints are tightly closed by effects of depth or
horizontal stress, VF will more closely approach the value
of VL without RQD being affected. Therefore whether
RQD, or the Bieniawski, 1989 RMR, or the Barton et al.,
1974 Q-value are used to correlate with seismic velocities,
a depth or stress correction is required for use below the
superficial penetration of surface refraction measure-
ments, in other words when depths exceed some few tens (a)
of metres. In fact a depth correction may be needed at
shallower depths, but with RQD, RMR and Q typically
increasing rapidly in the first tens of meters, a reliable
depth correction is problematic, since three variables
are changing at once (quality, depth, velocity).
Turk and Dearman, 1986, proposed a seismic fissuriza-
tion index K that was based on the difference between
P-wave velocity of a dry, intact sample loaded to 1/2 c
(half the uniaxial strength), and the velocity of the dry
rock mass.

V /2  Vmass,dry
K  (5.2)
Vmass,dry (b)

They showed that K was sensitive to increased porosity Figure 5.1 Fissuration index K in relation to in situ velocity for dry
rock masses. Turk and Dearman, 1986, with andesite
caused by weathering (e.g., for fresh or weathered andes-
data from above the water table, from King et al., 1978.
ites: n  1.9%, K  0.21, and for n  9.1%, K  0.68
respectively). When they analysed data from King et al.,
1978, concerning Vp and joint frequency measurements
from above the water table in andesite, K was shown to One of the most thorough analyses of seismic refrac-
vary from about 0.1 to 0.3 with increased joint frequency, tion measurements in mostly hard, jointed rock envi-
while the in situ velocity (for unsaturated conditions) var- ronments was that given by Sjøgren et al., 1979 and
ied from about 6 km/s to 5 km/s. This data and King Sjøgren, 1984. The authors’ experience from some
et al., 1978, source data are shown in Figure 5.1a and b. 113 km of P-wave surveys (15 sites) and 5 km of S-wave
Karmis et al., 1984, also investigated the effect of surveys (5 sites) were compared with the results of
fractures (saw-cut or tensile) on the seismic velocity. 2.85 km of core from 74 drill holes at 8 of the hard rock
When F(m1) was plotted against the velocity ratio sites. The range of rocks occurring at the measured
(V/V0), a linear relationship was given. In approximate locations, mostly in Norway, were: amphibolite, gran-
terms, the following was found: ite, gneiss, meta-anorthosite, pegmatite, porphyry,
quartzite and mylonite.
Fm1 V/V0% The authors were careful to emphasise that the cor-
relations they derived between P-wave velocity and joint-
18 50
ing descriptions such as mean RQD and mean frequency
9 75
F(m1) were relevant only to unweathered igneous and
4.5 90
metamorphic rocks, and generally for the upper 20 to
Relationships between rock quality, depth and seismic velocity 71

Figure 5.3 Mean values of physical and dynamic properties for hard,
unweathered igneous and metamorphic rocks, based
on shallow refraction seismics. Sjøgren et al., 1979.

close to the minimum, when considering a line sample


through the rock mass.
The general joint frequency and RQD trends for
these unweathered jointed rocks, including resulting
dynamic moduli are summarised in Figure 5.3. (In this
figure k  bulk modulus, and
 shear modulus.
This data is reproduced at larger scale in a subsequent
comparison with Q-values.)
Sjøgren et al., 1979, discussed various factors that
could alter the proposed mean joint frequency and mean
RQD versus Vp trends shown in Figures 5.2 and 5.3. They
pointed out that the ‘natural’ velocity of the unjointed (or
most massive) rock from site to site could vary due to rock
type, mineralogy, etc. (One could also add to this list the
inter-related technical terms: porosity, density and uni-
axial compressive strength.) Besides these fundamental
Figure 5.2 Curve 1  joints per metre, curve 2  RQD, as a causes for variation, the effects of weathering and depth
function of Vp, from shallow refraction seismic in hard, of measurement were obviously of particular influence.
unweathered rocks, mostly from Norway. Sjøgren et al., For this reason the authors addressed most of their
1979. attention and derived most of their data from the depth
zone of up to 20 or 30 metres. When they conducted
30 metres. They also emphasised the differences caused subsequent tunnel measurements, they found that a 30 to
by weathering. 50 m depth resulted in a general increase in velocity of
Figure 5.2 shows mean numbers of cracks (joints) about 5 to 15%, greatest for the lower velocity. When
per metre (Fm) for a given velocity in more or less Vp was lower than 3 km/s they had observed ‘consider-
unweathered Scandinavian rocks. The general trend for ably greater’ increases with depth, and also a common
mean RQD values versus velocity are also given for this reduction of the widths of the low velocity zones with
‘unweathered’ data set. Deviation from these average increased depth (40 to 60% was quoted).
curves were reported to be about 1 crack/m at the In a later publication, Sjøgren, 1984, gave his earlier
higher velocities, and 1.5–2.0 cracks/m for the lower example of hard rock correlations between mean RQD,
velocities. Corresponding dispersions of RQD values mean joint frequency per metre (Fm) and mean
were 2–3% and 5–6% respectively. P-wave velocity (shown in Figure 5.2) with an additional
The 74 drill holes were as close as possible to, or on the curve 3 related to the mean trend of RQD in Permian
seismic lines, and directed to be as perpendicular as pos- and Triassic sandstones (Figure 5.4). Obviously these
sible to the tectonic structure or foliation. Fm1 values sets of measured data cannot all fit the simple relation
are therefore close to the maximum and RQD values of Deere et al., 1967, that (V F/V L)2  RQD/100.
72 Rock quality, seismic velocity, attenuation and anisotropy

Figure 5.5 Data from Sjøgren and co-workers for Fm1 versus
Vp for various rock conditions, with increased weather-
ing effect and/or reduced strength, from right (#1) to
left (#4). Palmström, 1996.

Figure 5.4 Mean RQD and Fm1 as a function of Vp for the pre- the Andes and in Tanzania. In relation to empirical cor-
vious hard rock sites (curve 1  Fm1, curve 2  relation possibilities, corrections for weathering, poros-
RQD), and for Permian and Triassic sandstones. ity and rock strength (or density) are probably needed
Sjøgren, 1984.
to explain the range of data.
Extensive fracture frequency and velocity data were
presented by Niini and Manunen, 1970. The data in
The large reductions in velocity (1.5–2 km/s) are Figure 5.6 were derived from 55 vertical or steeply
clearly a function of increased porosity (and density) inclined holes drilled into the upper 15 metres of
and uniaxial strength (or deformation modulus E). bedrock, along 100 km of seismic traces made for the
When there is a tendency for weathering, or for matrix 120 km long Helsinki water supply tunnel.
porosities higher than normal for hard rocks, then the The possible complication of increased stress, from
same joint frequency will be recorded at lower seismic veloc- tectonic causes or from 15 to 30 m of additional soil
ities. The four curves from Sjøgren and co-workers, cover, has apparently meant that high fracture frequen-
shown in Figure 5.5 represent at the one extreme (curve cies were recorded even when velocities were as high as
No. 1) the same data as given in Figure 5.2 for hard, 4.5 km/s. Possibly all fractures in the core, and the natu-
unweathered rocks at shallow depth. The degree of ral joints, were each included in ‘fractures/m’, since
weathering increases, rock strength eventually reduces, these numbers for Fm1 are significantly larger than the
and the matrix porosity increases while progressing from Sjøgren data sets. Fracture frequencies were unusually
curve No. 1 to curve No. 4 in Figure 5.5. high (F  26/m for 0–5 m rock depth, F  21/m for
The data were assembled from Sjøgren and co-workers, 5–10 m rock depth) and seismic velocities were strongly
and are derived from measurements in Scandinavia, in affected as a result.
Relationships between rock quality, depth and seismic velocity 73

Figure 5.7 The separation of velocities within fractured zones, and


Figure 5.6 Joint frequency and velocity trends from the Helsinki outside the fractured zones, from studies for the 100 km
water tunnel. Niini and Manunen, 1970. There appears long Helsinki water tunnel. Niini and Manunen, 1970.
the possibility that both joints and artificial fractures
were counted, in view of the extremely high Fm1 magni-
tudes, for a given velocity.

The authors gave the following ranges of results for


intact rock samples obtained from 31 drillholes in the
granites and from 2 drillholes in the mica gneisses.

Vp (field) km/s Vp (lab) km/s Edyn (lab) GPa

Granites 3.4–5.3 5.8–5.5 61–82


Mica gneisses 3.8–5.2 5.5–5.8 85–96

An unusual set of data for fractured zone widths was


also described by Niini and Manunen, 1970. Zone
widths were shown in relation to velocities measured
within the low velocity zones and also outside the
zones. Figure 5.7 shows that fracture zone widths were
larger, as one would expect, if the velocity outside the
zone was also low. Narrowest zones tended to have low-
est internal velocities, and highest external velocities. Figure 5.8 Examples of joint frequency effects on downhole acoustic
Developments in logging joint and fracture frequency log Vp values, with comparison to laboratory velocities
effects using a downhole acoustic tool that could log in under the same stress levels as in situ. King et al., 1978.
water-filled or dry boreholes were described by King
et al., 1978. Figure 5.8a and b show good examples of the
relationship between the larger scale downhole velocities The highly fractured and altered zones were, of course,
and the laboratory velocities from a mining location in strongly correlated to velocity and amplitude reductions.
andesites and pegmatites. Laboratory conditions of The same authors also made some useful assessments of
humidity and stress (7 MPa) were matched to the min- the effect of logged fracture (or joint) frequency (Fm1)
ing stope conditions as far as possible. The data shown on velocities in cored 60 m deep boreholes (Figure 5.8),
in Figure 5.8 can also be taken as a useful example of and assembled other near-surface and underground data
excavation disturbed zone (EDZ) and drainage effects, to investigate correlation of the squared velocity ratio
although since joint frequency also increases towards (VF/VL)2 with joint frequency (Fm1). Care was taken to
the mine opening, the combined effects of Fm1 and discount the drilling induced fractures. The authors gave
% saturation cannot be separated. the following correlation (see Figure 5.9).
74 Rock quality, seismic velocity, attenuation and anisotropy

and Q-logs including a useful set from the Xiaolangdi


multi-purpose hydroproject in China, where plate load-
ing tests also provided ‘static’ deformation modulus
measurements that showed broad correlation with the
Q-values independently logged by the writer.
Various hard rocks such as granites, gneisses, volcanic
ignimbrite and competent sandstones were also tested in
projects in Norway, England and Hong Kong, where
first-hand information on Q-logging had been obtained.
The proposed relationship, which was briefly intro-
duced in Chapter 1, was as follows:

Vp  3.5  log10 Q (5.4)


Figure 5.9 A collection of near-surface and underground data,
showing velocity ratio squared (VF/VL)2 as a function This empirical relation is plotted in two different
of fracture frequency. King et al., 1978. ways together with Sjøgren et al., 1979, trends for RQD
and Fm1 (derived from Figure 5.2). Figures 5.10a and b
show the result. When the rock quality Q-value is 1.0,
(VF /VL )2  0.96  0.036F (5.3) midway between the extremes of 0.001 and 1000,
Vp  3.5 km/s. The velocity changes by roughly 1 km/s
(Correlation coefficient  0.93) (upwards or downwards) for each ten-fold change in
rock quality Q-value.
This model has now been tested on sites in several
5.2 Relationship between rock countries where rock quality Q-logging of core has
quality Q and Vp for hard jointed, been performed. The fit to measured data is quite good,
near-surface rock masses provided that depths are shallow (i.e. down to 25 m,
near the usual limit of shallow refraction seismic surveys
Due to the seismic ‘visibility’ of jointing in the upper 25 to performed where depth of weathering is relatively lim-
30 metres, Sjøgren et al., 1979, and Sjøgren, 1984, were ited). A further necessary condition is that the rocks are
able, as we have seen, to record significant correlations non-porous and reasonably hard (i.e. typically with
between Vp, RQD and joint frequency. Since their meas- uniaxial strengths of 100 MPa or more). This model for
urements were shallow, the effect of stress-induced joint hard rocks, and a modified one for soft porous rocks to
closure was minimised. They also effectively removed be developed later, can be used for initial interpretation
other sets of variables by generally recording correlations of seismic data.
for hard and almost unweathered igneous and metamor- The table of data given in Figure 5.11, from Sjøgren
phic rocks. The variables of depth, porosity, uniaxial et al., 1979, can also be expanded to include the
compressive strength and density were therefore largely Q-value scale, as shown at the bottom of this figure. As
removed. This has some advantages for what follows. also noted in the figure, depth or stress effects, discussed
A hard rock, near surface correlation of Vp and rock in detail later, will mean that the Q-scale must be shifted
quality Q-value can also be derived on the same basis as more and more to the right in relation to Vp, as depth
above, if effects of porosity, uniaxial strength and depth increases. The same of course will actually apply to the
are first ignored, as for the Sjøgren et al., 1979, data RQD and Fm1 scales. For the above reasons, the sug-
shown in Figures 5.2, 5.3 and 5.4. gested correlations must be strictly applied to near-
On the basis of the Norwegian Geotechnical Insti- surface seismic data (shallow refraction seismic or shallow
tute’s cross-hole seismic tomography measurements at boreholes) in hard, non-porous, largely unweathered
the Gjøvik (62 m span) cavern site in Norway, shown in rocks, but of course can apply to heavily fractured and
Chapter 4, and based on Q-logging of the same bore- sheared zones, (i.e., faults) containing clay.
holes, a preliminary model for a Vp-Q-value relationship It is interesting to note the perceived inter-relationships
was developed by Barton, 1991. This was subsequently between eight methods of rock mass classification using
confirmed by analysis of several other seismic, cross-hole the Chinese descriptive classes ‘soft rock’, ‘hardpan’,
Relationships between rock quality, depth and seismic velocity 75

Figure 5.10 a) Mean RQD and Fm1 trends for hard, near-surface, low porosity rock masses, from Figure 5.2 Sjøgren et al., 1979, with an esti-
mated extrapolation of the ‘extremes’ (dashed-lines), by the writer. The lower rock quality Q-scale, added by the writer, is also only
relevant for hard, unweathered, low porosity, near-surface rock masses. b) Note rearrangement of Q-Vp relationship, with appended
RQD and Fm1 data from Sjøgren et al., 1979, after Barton, 1995.

Figure 5.11 Mean RQD, Fm1, Vp and Edyn. data for hard, near-surface, low porosity rock masses, from Sjøgren et al., 1979. The rock qual-
ity Q-value scale was added by the writer. Note the need for a shift of the Q-scale to the right, with increasing depth.

‘normal soil’ etc. that was given by Chen, 1982, using the Comparison of rock quality Q-values logged in bore-
seismic velocity ranges as a reference. Approximate simi- holes (or mapped at the surface) and seismic velocity
larity to the empirical model (Vp  3.5  log Q) that measurements are not yet very common in the rock
was developed ten years later is indicated in Figure 5.12. mechanics literature, though data is available at numerous
76 Rock quality, seismic velocity, attenuation and anisotropy

Figure 5.12 Inter-relationships between various rock mass classification schemes, Vp, RQD, and the rock quality Q-value. Chen, 1982.

sites. Chan, 1993, describes the engineering geological


investigations performed for the purpose of optimising
the location and orientation of a 24 metre span cavern
to be used for refuse transfer, at Mount Davis in
Hong Kong.
General mapping of tuffaceous outcrops using the
rock quality Q-system (70% Q  10, 25% Q  10 to
2, 5% Q  2 to 0.1) was supplemented by inclined
drill holes and surface refraction seismic measurements.
Figure 5.13 compares the results of core-logging (Q
range 1 to 100) and an adjacent seismic profile,
showing near-surface velocities of 4.5, 4.7 and 5.9 km/s at
about 25 m depth.
Velocities of only 1.1 and 3.2 km/s were recorded in a
weakness zone. Based strictly on the simple Q–Vp rela-
tionship for low porosity, hard rocks at shallow depth
(Vp  3.5  log10 Q), the following wide range of
Q-values can be anticipated from the above velocities: Figure 5.13 Comparison of refraction seismic velocities and core-
logged rock quality Q-values at a cavern site in Hong
Vp  5.9, Q  250 very good quality Kong tuffs. Chan, 1993.
Vp  4.7, Q  16 good quality
Vp  4.5, Q  10 good quality
Vp  3.2, Q  0.5 very poor quality zone shown in Figure 5.13 is clearly quite complex at
Vp  1.1, Q  0.004 exceptionally poor quality cavern level (50 to 70 m depth) since it has rock quality
Q-values of less than 1 (two locations), and Q-values
These appear to be reasonable values for massive rock, of about 10, 20 and 70 in other parts of the zone. The
good quality rock, and fault zones respectively. How- shallow seismic picks up Vp  1.1 (Q  0.004?) and
ever, a basic problem with such comparisons is that Vp  3.2 (Q  0.5?) but not the higher quality slabs (?)
the shallow refraction seismic is not penetrating to the of rock apparently existing at greater depth within the
same depths as the core-logging. The broad, low velocity same fault zone.
Relationships between rock quality, depth and seismic velocity 77

5.3 Effects of depth or stress on velocity in a weak porous rock. This was due to some for-
acoustic joint closure, velocities tuitous circumstances at a test tunnel in chalk. Their stud-
and amplitudes ies were performed in an experimental machine-bored
tunnel in the Lower Chalk at Chinnor in Oxfordshire,
Bertracchi et al., 1966, gave some early Italian experi- England, and also in the laboratory. All the studies were
ences of cross-hole and downhole sonic logging, and performed in chalk from above the water table, but with
noted a consistent tendency for increased velocity with a natural water content of about 17 to 20% and a density
depth (usually 30 to 60 m/s increase per metre) in the of 2.2 gm/cm3. Firstly, Figure 5.15a shows the influence
depth range 5 to 25 m. However, related core logging of intact strength on Vp values, using penetration tests at
results were not given. This increase in velocity gives an ch. 79 m in the tunnel. A similarly strong effect of joint or
extreme gradient of 30 to 60 s1, which is about an discontinuity spacing on Vp, from the same location in
order of magnitude greater than in the subsequent few the tunnel is shown in Figure 5.15b.
hundred meters. The most interesting result was the effect of stress
Depth effects on seismic velocity were also reported level (or tunnel depth) on seismic velocity. Velocity
by Cecil, 1971, from a survey of Swedish tunnels. Velo- increased from typical values of 1.1–1.3 km/s in the
cities at 50 to 60 m depth in high quality rock were up first 30 metres of tunnel, up to 1.5–1.6 km/s between
to 17% higher, while for low quality, heavily jointed 40 and 60 m inside the tunnel, where overburden had
rock they were up to 38% higher. By comparison Sjøgren increased to some 15–20 m. This increase in Vp might
et al., 1979, reported 5 to 15% increase at 30 to 50 m appear to be as expected, but it actually occurred
depth compared to that at the surface. In both the above despite an increase in the frequency of joints and dis-
cases the comparative rock qualities at the different depths continuities in the chalk, as one progressed further into
are a factor of uncertainty. the tunnel (Figure 5.16). Often, Vp-depth data can be
There are also cases in the literature where depth ambiguous because velocity increases occur at depth,
effects are, apparently, absent. Bertacchi and Sampaola, due also to less frequent jointing. Here the two effects
1970, conducted repeated measurements of seismic were, by unusually good fortune, separated.
velocity at four Italian dam sites, using a combination of New and West, 1980 working on the same problems,
downhole sonic logging and cross-hole measurements. also performed loading tests on artificially fractured or
The deepest measurements (to 100 m depth) were con- sawn interfaces for various rocks, and found that for sev-
ducted at the 160 m high Alpe Gera dam, founded in eral different surface roughnesses in the case of the chalk
serpentine with marked foliation. Measurements were from the Chinnor Tunnel, a stress of about 0.4 MPa was
repeated over a four-year period and showed invariance required for ‘acoustic closure’ (Figure 5.17). Signifi-
with time, and independence of reservoir storage level. cantly, this stress also corresponded to the overburden
In view of the dependence of Vp on effective stress, it is stress where in situ Vp values had shown a certain flat-
tempting to assume that the measurements were con- tening out. The maximum in situ Vp values of about
centrated downstream of a successful grout curtain and 1.6 km/s (influenced by a fracture frequency as high as
effective drainage. Entirely different experiences are seen ten per meter), may be compared with Vp values for
at other dam sites, as we shall see in a later chapter. intact blocks of 1.95 km/s (range 1.6 to 2.2 km/s).
An illustrative set of depth-related data, which never- Before leaving the Chinnor Tunnel, it may be of inter-
theless is inconclusive, is that given by Mouraz Miranda est that Hudson, Jones and New, 1980, also mentioned
and Mello Mendes, 1987, in Figure 5.14. The objective very low P-wave velocities (0.6–1.0 km/s) for badly frac-
was to explain the reducing rates of penetration and tured areas of the chalk, and quoted permeability values
increased specific energy used when diamond core drilling of 106 to 104 m/s, or approximately 10–1000 Lugeons.
and downhole hammer drilling in a 22 metre deep profile According to a rule-of-thumb (L  1/Qc developed in
of weathered granites. Since all the indices of quality used Chapter 9), strength-normalized rock quality Qc -values
were increasing with depth (i.e. hardness, RQD, density), might then be expected to range from about 0.1 to 0.001,
it is inevitable that a very large increase in Vp (1.0 to in the absence of complications caused by clay sealing.
4.5 km/s) should have been registered. This range of qualities (where Qc  Q  c/100) is
Hudson et al., 1980 (‘rock’ Hudson), demonstrated broadly what might be expected from rock quality
some fundamental effects concerning the influence of Q-logging in badly fractured areas of this weak rock,
rock strength, joint spacing and depth (or stress) on seismic since if one assumes values of uniaxial compressive
78 Rock quality, seismic velocity, attenuation and anisotropy

Figure 5.14 Simultaneous increases in hardness, RQD, and density give a cumulative effect on Vp (450%) and drilling rate (75%).
Mouraz Miranda and Mello Mendes, 1987.

strength ( c) for the chalk from about 10 MPa down to of this chapter), since they argued that stress could
1 MPa (when weathered), the implied rock quality ‘acoustically close’ joints, and joint frequency as expressed
Q-values would be respectively 1 and 0.1, closely by RQD would then prove to have little effect on the
resembling ‘well-jointed’, and ‘heavily jointed and velocity. Their experiments on artificial flat surfaces in
weathered’ rock, respectively. contact, to simulate smooth joints in various rocks, show
New and West, 1980, also discussed the question of the approach of ‘acoustic closure’ at about 2 to 4 (?) MPa
stress level (or tunnel depth) on joint closure. They normal stress in Figure 5.17. However, these artificial
questioned the applicability of the Deere et al., 1967 rela- surfaces were probably not responding quite in the usual
tion (i.e., (Vp field/Vp lab)2  RQD/100: see beginning non-linear manner, in terms of stress-closure, such as
Relationships between rock quality, depth and seismic velocity 79

Figure 5.16 Seismic measurements at the Chinnor Tunnel in


England. Increase of overburden stress caused V p to
increase (as expected), but this was despite an unusual
increase in joint frequency with greater depth. ‘Absolute
proof’ of a stress-related increase in Vp is evident from
Figure 5.15 Penetration strength and discontinuity spacing show this case, since the matrix did not change. Hudson,
a strong influence on the P-wave velocity for partly Jones and New, 1980.
saturated, porous chalk from the Chinnor Tunnel.
Hudson, Jones and New, 1980.
stress, as shown by the Chinnor experience, but large
stress increases will inevitably ‘reactivate’ sensitivity to
reported by Bandis et al., 1983, and they only represented stress as compaction occurs.
the behaviour of ‘joints’ that were perpendicular to the ‘Acoustic closure’ studies by Westerman et al., 1982,
principal stress. were based on artificially sawn, ground and then acid-
In Figure 5.18, New and West, 1980, show the etched surfaces of a corallian limestone. Perhaps because
insensitivity of Vp to joint frequency changes in a sand- of the partial ‘non-mating’ of the etched depressions in
stone, as measured along the Kielder aqueduct tunnel, each surface they found that normal stress levels as high
in northern England. However the principal joints were as 20 MPa were required to reach the seismic velocity
reportedly vertical only (i.e. one set that had responded (5.6 km/s) of the intact rock. The parabolic-type increase
to the effects of 100 m overburden and previous geo- in velocity from 3.9 km/s was quite smooth, and corres-
logical history, and were perhaps tight and closed). ponded to a reducing attenuation, the latter stabilising
In a mudstone section of the same tunnel, calcite beyond normal stresses of about 10 MPa. The samples
filled joints and some heavily jointed and faulted zones were dry as far as can be understood from the results
showed generally strong effects on velocity, as seen in obtained.
Figure 5.19. In general, a rock mass with several joint The ratio of stress to strength, for example the ratio
sets is likely to show stress sensitivity to greater depth or JCS/ n used in the shear strength criterion of Barton and
stress level than the authors are implying, and if the Choubey, 1977 may be useful for explaining so-called
rock is very weak and porous, volume changes will ‘acoustic closure’, since it is not only the stress level, or
occur at greater stress giving velocity increases. In a rock depth, but also the rock joint stiffness or strength that
like chalk marl, joints may close easily with moderate determines the contact area needed for the less attenuating
80 Rock quality, seismic velocity, attenuation and anisotropy

Figure 5.18 Insensitivity of Vp to vertical joint frequency in sand-


stone at the Kielder Tunnel, with 100 m overburden.
New and West, 1980.

Figure 5.17 Vp increases for flat, dry artificial rock surfaces under
normal stress. New and West, 1980.

seismic transmission across joints. The ratio of JCS


(joint wall compression strength and n (effective
normal stress) is closely related – or perhaps nearly
identical – to the ratio of A1 (assumed contact area)/A0
(actual contact area). The ratio A0/A1 is very small for
the case of hard unweathered rock joints near the sur- Figure 5.19 Sensitivity of Vp to frequencies of calcite-filled joints
face, which continue to show velocity increase for many in mudstone section of Kielder Tunnel. New and
hundreds of metres, while A0/A1 is much larger for the West, 1980.
case of a soft porous rock mass like chalk marl, which
can show ‘acoustic closure’ at stresses as low as 0.4 MPa
according to Hudson, Jones and New, 1980. below this aperture. The above effects and new experi-
Tanimoto and Ikeda, 1983, found that Vp was approx- mental data are the main reason why Tanimoto and
imately proportional to the normal stress applied to Kishida, 1994, and others are advocating the use of com-
simulated joints over the range 3 to 20 MPa, but that Vp pression wave amplitude as a supplement to velocity
dropped rapidly below stress levels of 3 MPa. A cut-off data, for better sensitivity to jointing, particularly for the
aperture of about 40
m separated their experimental higher stress levels than those usually employed in shal-
results, with apparently no influence of Fm1 on Vp low refraction seismic.
Relationships between rock quality, depth and seismic velocity 81

Figure 5.20 Tests on artificially fractured columns of marble,


showing Vp and Vs as a function of normal stress, in
the dry state. Stacey, 1977.

Laboratory tests using artificially jointed columns of


marble that could be axially loaded and submerged in
water were described by Stacey, 1977. These very inter-
esting and instructive tests showed that the commonly
used P-wave velocity was not the most sensitive param-
eter as regards rock quality. ‘Rock quality’ in this set of
experiments was simulated by having 0, 1, 2 or up to
8 tension fractures distributed along the 0.46 m length
of marble under dry or wet conditions, or with clay in
the fractures. The basic P-wave and S-wave velocity
responses to different levels of normal loading (0 to
7 MPa) in the dry state are shown in Figure 5.20.
The number of fractures corresponds to F values
(m1) of about 2, 6, 9, 13 and 17 per metre. The major
increases in P-wave velocity (e.g., 2.8 to 4.4 km/s) occur Figure 5.21 Tests on artificially fractured columns of marble. a)
in the first 2 MPa of normal loading, thereafter less S-wave frequency. b) relative P-wave amplitude. Stacey,
rapid increases are seen, but the rise is consistent and 1977.
nearly constant up to the maximum applied stress of
7 MPa. The number of fractures had much less influ- Stacey, 1977, went on to demonstrate that shear wave
ence on velocities at stresses above 2 to 3 MPa, in frequency was also very sensitive to the degree of joint
the dry state. Under zero stress levels the number of closure caused by stress. Frequency increases from
fractures had the greatest influence. When, by contrast, about 8 kHz to about 19 kHz were indicated for the
the tension fractures were wet and clay-coated, their case of well compressed, dry tension fractures, with less
number reportedly had ‘negligible’ effect on P-wave sensitivity in the case of wet conditions, or with wet
velocity. clay fillings. These results are shown in Figure 5.21a.
82 Rock quality, seismic velocity, attenuation and anisotropy

(There appears to be a possible link here to shear wave


splitting, polarization and anisotropy interpretation,
for the case of fractured reservoirs with gas or liquid sat-
uration – see Chapters 13 and 15.)
Furthermore, the compressional wave amplitude showed
great sensitivity to both stress level and the frequency of
the tension fractures, as demonstrated in Figure 5.21b.
Compressional wave amplitude is of course an indicator
of the level of attenuation, which was shown to be maxi-
mum in the case of unloaded, multiple-fracture models,
and minimum for the case of well confined, single fractures.
Attenuation is treated in detail in Part II, Chapter 10.
A related though more comprehensive set of experi-
ments was reported by Tanimoto and Ikeda, 1983,
using numerous 50 mm diameter cylinders of various
rocks (rhyolite, sandstone, granite, granodiorite and
tuff ). The rock cylinders had five different lengths,
making it possible to represent a ‘line sample’ through
and imagined rock mass, with 0, 10, 20, 30 or 40 joints Figure 5.22 Schematic of the multi-cylinder ‘line samples’, indi-
per metre. Filter paper, dry or saturated, was used to vidual ‘filled-joints’ and a natural joint, that were
produce the desired ‘joint apertures’ and moisture con- studied by Tanimoto and Ikeda, 1983.
tents. Figure 5.22 shows a schematic of the various ‘line
samples’ created by cylinders of different length. In the
case of the natural joint sample, both normal and shear
load was applied.
In Figure 5.23, the authors showed that the ratio
k  Vp (jointed)/Vp (intact) was hardly influenced by
the frequency or number of the types of ‘joints’. How-
ever, they qualified this by indicating that it was their
apertures of less than 0.04 mm (or 40
m) that caused
the frequency of fractures to have little influence on the
P-wave velocity. In the context of numerous seismic
surveys underground that they had included in their
evaluations, they suggested that apertures were consid-
erably wider than 40
m, therefore allowing fracture or
joint frequency (Fm1) to influence the velocity ratio
VJ/Vi as follows:

1
 5  2

Vjointed  Vintact   (5.5)


 F  4  Figure 5.23 Joint frequency (n), velocity ratio (k), and amplitude
ratio A/A0, indicating little influence of ‘joint’ fre-
quency (or Fm1) on the P-wave velocities, but a strong
No reduction in P-wave velocity is predicted with 1 influence on the amplitude ratio. Tanimoto and Ikeda,
joint per meter, but with 5, 10 or 20 per meter, the ratio 1983.
k is predicted to show successive reductions to 0.75, 0.6
and 0.5 (to the nearest decimal places). So a hard crys-
talline rock with Vintact of 5.5 km/s, might show reduc- One may perhaps conclude that heavily jointed rock
tions to 4,100, 3,300 and 2,500 km/s respectively. This masses have a degree of acoustic coupling that is not
appears inherently representative of experiences in frac- as complete as in the (aperture 40
m) laboratory
tured zones, at least in the near-surface. experiments, due to variously oriented joints and lower
Relationships between rock quality, depth and seismic velocity 83

near-surface stresses, various rock-to-rock contact areas, Fratta and Santamarina, 2002, also used columns of
and various ranges of weathering, mineral coatings and blocks under stress to demonstrate velocity-stress sensitiv-
fillings. The contact area ratio A0/A1 referred to earlier ity. They studied the effects of varying thicknesses of kao-
in this chapter will obviously change completely in the linite gouge materials, finding that shear wave velocity
presence of clay filling under high compressive stresses. gave a very sensitive indication of the effect of gouge thick-
As in the case of Stacey, 1977, Tanimoto and Ikeda, ness at even lower stress levels, equivalent to about 1 to
1983 found that compressive wave amplitude gave a very 10 m, typical of the weathered zone. For example at
sensitive measurement of the number of ‘joints’. The 0.25 MPa normal stress, equivalent to about 10 m depth,
amplitude ratio A/A0 (A0 for intact rock) is shown in the conditions a) no gouge, b) 0.5–1 mm of gouge, c)
Figure 5.23 (open circles), and shows excellent sensitivity 2.0 mm of gouge, d) 2.5 mm of gouge, gave S-wave veloc-
to number of ‘joints’. These authors also found that aper- ities of about 850 m/s, 750 m/s, 600 m/s and 450 m/s
tures less than 0.01 mm (10 ␮m) did not have influence on respectively. The strongly non-linear (convex) Vs – normal
the wave propagation, even when the normal stress was as stress curves, showed velocities ranging, respectively, as low
low as 1–2 MPa. Physical apertures (E) of this size (in as 450 to 250 m/s at an equivalent depth of only 1 meter.
contrast to hydraulic apertures (e) which are smaller due In hard rocks, such as the numerous cases reported by
to roughness effects), are probably rare in the upper Sjøgren et al., 1979 and Sjøgren, 1984, there is a signifi-
20–30 metres of rock masses where refraction seismic is cant in situ correlation between Vp and joint frequency
carried out, so this result is probably consistent with (and RQD), due presumably to the fact that this ‘cut-off
experience in the field. aperture’ (whatever it may be in different rocks) has not
(One may speculate whether such a finding could been reached at the moderate (20–30 metres) penetra-
have application in the interpretation of so-called ‘open’ tion of shallow refraction seismic surveys. Tanimoto and
fractures in petroleum reservoirs: would these need to be Ikeda, 1983, found that Vp was proportional to n in
of greater aperture than this order-of-magnitude, before the stress range 3–20 MPa, but dropped sharply for
they could cause shear wave splitting?) n  3 MPa. By chance, or similar physics, in the field
A combination of permeability testing and index case records used to design the empirical Vp–Q–M chart
testing of the relevant joints (i.e. roughness JRC, and to be shown later (Barton, 1995), Vp and depth are also
wall strength JCS) using the methods described by found to be linearly related from about 200 m to
Barton et al., 1985, for converting between hydraulic 1000 m depth (5 to 25 MPa), with Vp falling rapidly for
and physical joint apertures would indicate the rough depths in the range 25 m to 100 m, i.e., for n or v
order of magnitude of the ‘necessary’ hydraulic aper- (or h) 2.5 MPa.
tures to satisfy this possible ‘rule-of-thumb’ that E must
be 10
m, for having influence on wave propagation.
A look ahead to Chapters 15 and 16 where joint 5.3.1 Compression wave amplitude
properties are discussed in detail, would suggest that sensitivities to jointing
hydraulic apertures of about 10, 1.8 and 0.3
m might
be operating with E  10
m, if joint roughnesses One of the most thorough studies of the effect of joint
were respectively 2.5 (quite smooth and nearly planar), parameters on seismic signatures was reported in a sub-
5 (near-planar but some small undulations) and 10 sequent study by Tanimoto and Kishida, 1994 and
(non-planar with marked inclined asperities). These lat- Kishida, 1999, which was built on these earlier investi-
ter would hardly be considered as ‘open’ joints, and in gations of Tanimoto and Ikeda, 1983. The advantages of
a reservoir situation are perhaps (as suggested by the compressive wave amplitude compared to Vp for sensing
Tanimoto and Ikeda, 1983 results) not capable of influ- joint frequencies at the higher stress levels was empha-
encing seismic waves, nor, by implication, shear wave sised again, and convincingly demonstrated experimen-
splitting. tally. The authors also compared (conventional) seismic
Tanimoto and Ikeda, 1983, also investigated the velocity tomography (SVT) with seismic amplitude
effect of larger apertures (or thicknesses of discontinu- tomography (SAT). The latter reportedly corresponded
ity fillings) using more sheets of filter paper to give a more closely to jointing observed with a borehole scan-
range of 1 to 3.4 mm thickness. There was strong sensi- ner, as compared to the more frequently used (SVT).
tivity of Vp to stress level (0.3 to 3 MPa) and to the total The studies were made in boreholes at a dam site, using
cumulative joint aperture and moisture condition. an exploratory adit for further confirmation.
84 Rock quality, seismic velocity, attenuation and anisotropy

The extensive laboratory studies reported by


Tanimoto and Kishida, 1994, were based on cylindrical
rock specimens, with a total of 86 cylinders represent-
ing sandstone, rhyolite, shale, granite, granite porphyry
and slate. In the case of ‘joint samples’, natural joints
were moulded and replicated with hard mortar or plas-
ter, to reproduce the same roughness as in nature. Since
the authors were very concerned about the correct meas-
urement of roughness they utilised a non-contact laser
scanner micrometer mounted on a 3D, movable stage
roughness profiler.
They show in Figure 5.24 a relationship between
the joint roughness coefficient JRC (from Barton and
Choubey, 1977) and ‘Ms’. The latter is derived from a
power spectrum versus frequency relation analysed by Figure 5.24 Correlation of the parameter Ms used by the authors,
the so-called ‘maximum entropy method’, which the with JRC for the natural joints. Tanimoto and
Kishida, 1994.
authors selected in preference to the ‘fast Fourier trans-
form’ method conventionally used. The inter-relationship
between ‘Ms’ and JRC shown in Figure 5.24 is needed amplitude as expected, and the opposite occurred (loss
to interpret the effect of joint roughness variations used of energy), in the case of shearing or joint dilation ‘post
in their seismic velocity and seismic amplitude (A/A0 peak’.
results shown collectively in Figure 5.25 and 5.26. Each These comprehensive results leave one in no doubt
horizontal pair of figures compares the sensitivity of Vp about the potential advantages of amplitude measure-
and amplitude ratio A/A0 to the following: ment compared to velocity measurement. Even such
details as joint roughness, which obviously also relates
1. ‘joint’ frequency (25, 37.5 or 50 m1) and normal to acoustic coupling and contact area, demonstrates
stress level (0 to 3 MPa) that amplitude measurement has clear sensitivity to the
2. ‘joint’ aperture (0, 1.05 or 2.10 mm) and normal mechanics involved in the joint plane. The ratio A/A0
stress level (0 to 3 MPa) tends to get smaller (increased attenuation) as rough-
3. joint roughness (see Ms–JRC relation) and normal ness increases, which fits with the picture of joint clos-
stress level (0 to 3 MPa) using joint replicas ure difficulties when joints are rough. (See extensive
4. shear displacement of joint replicas (0 to 7 mm) for treatment of rock joint behaviour in Chapter 16).
three normal stresses (0.2 to 0.6 MPa) The above authors would probably be the first to
5. dilation (0 to 1 mm) during joint shear at three agree that further studies with real joints would be an
normal stresses (0.2 to 0.6 MPa) for JRC  13 advantage, and of course that further field studies using
6. aperture change during normal loading of joint velocity and amplitude tomography, together with com-
(JRC  8 to  13). prehensive joint surveys, would be necessary for refining
the interpretation of these useful techniques. Since SAT
These two figures demonstrate why shallow seismic and SVT methods reportedly show good correspon-
refraction measurements which operate at low stress dence when filling materials and wider apertures are
levels are successful in distinguishing joint frequency and found (i.e., near surface), an understanding of the inter-
aperture (achieved here with moist filter paper), but relationships can perhaps best be obtained in this low
why amplitude measurements appear to be much more stress, weathered, or partly weathered zone, since wide
desirable than velocity measurements at higher stress ranges of both Vp and (A/A0) are seen, and the rock
levels, if joint frequencies and character are still to be mass quality Q-value also varies strongly in this region.
distinguished. This emphasises the value of attenuation Some useful indicators about rock joint closure
measurement, or of seismic Q. It reflects richly of the mechanics can also be obtained from a study reported by
mechanisms involved with seismic detection of joints. Nihei and Cook, 1992. They utilised a combination of
Better ‘joint’ inter-locking from increased stress, acoustic emission (AE) monitoring and P- and S-wave
or an absence of dilation and shear, gave increased monitoring of artificial tension fractures in sandstone.
Relationships between rock quality, depth and seismic velocity 85

Figure 5.25 Parallel comparison (see a and b pairs) of Vp-monitoring, and amplitude ratio monitoring, of loading effects (0 to 3 MPa) with:
top) varied ‘joint’ frequency, centre) varied ‘joint’ filling thickness, bottom) varied joint roughness for the natural joints.
Tanimoto and Kishida, 1994.

These rough fractures were loaded to 24 MPa in mated The velocity-stress behaviour shown in Figure 5.28
and unmated conditions. As a result they produced indicated a consistent pattern of behaviour, with
widely different levels of acoustic emission (Figure 5.27), increased contact area at higher stress causing marked
especially on the first cycle of loading. increases in both the P-wave and S-wave velocities. The
86 Rock quality, seismic velocity, attenuation and anisotropy

Figure 5.26 Parallel comparison (see a and b pairs) of Vp-monitoring, and amplitude ratio monitoring of a) shear displacement along a nat-
ural joint, b) dilation during shear, c) aperture changes with three joints of different roughness. Tanimoto and Kishida, 1994.

mated fracture more closely approached the intact but at velocities of about 0.2 km/s slower, both for P-
sample in terms of velocity, showing rapid increases in waves and S-waves.
Vp from about 2.8 km/s (at 2.5 MPa) to 3.4 km/s (at It is apparent that rough-walled rock joints with
10 MPa), and thereafter a slower increase of velocity. their typically high JRC values (often 15 to 20) are
The unmated fracture showed nearly parallel behaviour, more difficult to close acoustically, so rough joints in
Relationships between rock quality, depth and seismic velocity 87

Figure 5.27 a) Stress-deformation cycles comparing intact, mated,


and unmated tension fractures. b) The number of Figure 5.28 A comparison of P-wave and S-wave velocities when
acoustic emission events associated with loading the loading the intact, mated and unmated samples from
mated and unmated fractures to 24 MPa. Nihei and about 2.5 to 25 MPa. Strong stress sensitivity is shown,
Cook, 1992. due particularly to the dry state. Nihei and Cook, 1992.

hard rocks could be seismically ‘visible’ to significant In massive granite at the Underground Research
depths (e.g., 10 MPa or more, corresponding to Laboratory (URL) in Manitoba, Canada, micro-seismic
several hundreds of metres overburden). There are EDZ and stress-induced failure sensing reported by
however other factors involved, such as mechanical Talebi and Young, 1992, showed P-wave velocities ran-
over-closure, and thermal over-closure, (Barton, 2004), ging from 5.6 to 5.9 km/s, and S-wave velocities ranging
which would give tighter apertures than ‘expected’ from 3.3 to 3.4 km/s for the depth range 310 to 400
from the present depth of burial or exposure. (See metres (approximately) down the 4.6 m diameter shaft.
Chapter 16). (The ratio of Vp(mean)/Vs(mean) was exactly 1.70 in this
The above closure aspects would contrast with the massive granite.)
evidence from some of the experimental work on flat Velocities increased by about 0.1 km/s for every
surfaces in weak rock reviewed earlier, for example that 30 m increase in depth, (gradient  3.3 s1), based on
of New and West, 1980, which would suggest much the 1 m deep measurements using numerous shallow
lower levels of stress sensitivity in the case of smoother boreholes drilled into the walls of the shaft. If stress-
rock joints, except when closer to the surface. induced fracturing had been involved, a linkage
88 Rock quality, seismic velocity, attenuation and anisotropy

between increased depth and increased velocity would


be less certain, as dilation and reduced velocity might
equally well occur, with such shallow holes. (Core recov-
ered from different depths at URL showed reduced
compressive strength and modulus with increased
sampling depth, due to increases in stress-induced
microcracking.)

5.3.2 Stress and velocity coupling


at the Gjøvik Cavern site

The foregoing review of depth and stress effects on P and


S-wave velocities measured across loaded joint assem-
blies, or at tunnel sites in natural settings, can be con-
cluded by a brief review of the stress effects on velocities
that were documented at the Gjøvik Olympic cavern site
in Norway. Figure 5.29 is a reminder of some of the
cross-hole seismic tomography discussed in Chapter 4.
Figure 5.30a shows the stress measurement results at
the cavern site, based on hydraulic fracturing and
hydraulic jacking of existing joints, performed by Tun-
bridge, of NGI. Both h min and the estimated H max
show rather high values, many times greater than the cal-
culated vertical stress gradient. The jointed gneiss at the Figure 5.29 Cross-hole seismic tomography at the Gjøvik cavern
site had steeply dipping conjugate jointing with up to site, showing strong increases in velocity with depth,
four joint sets, but fortunately these were joints with sig- actually caused most by high horizontal stress levels
nificant roughness and undulation. (3 to 6 MPa) in the upper 50 meters. There were veloc-
ity-depth gradients as high as 80 s1 in the upper
Based on the evidence shown in the velocity tomo-
20 m, and as high as 40 s1 over the first 50 m.
grams (Figure 5.29 and other tomograms from Chapter
Remarkably, the rock quality parameters: RQD, Fm1
4), a maximum velocity rise of some 2 km/s (3.5 to and Q, did not improve beyond about 5 m depth.
5.5 km/s) occurred in the depth range 10 to 60 metres, Barton et al., 1994.
over which range, horizontal stresses may have increased
by 2 MPa to 4 MPa, depending on direction relative to
H max and h min (Figure 5.30a).
The above rise of velocity occurred in a rock mass 5.4 Observations of effective
with a typical Q-value of about 8 to 10, with a joint fre- stress effects on velocities
quency that ranges from about 3 to 10 per metre, and
an RQD that ranges from about 60 to 100%. There is Dam construction represents a significant local source
no evidence from the four core logs of consistently of rock mass loading, in which the total vertical stress
increasing rock quality with depth, so the velocity initially is increased without necessarily changing the
increase of nearly 2 km/s occurred mostly as a pre- pore pressure, since the reservoir takes time to fill.
sumed result of the increase in horizontal stress. The When grouting of the foundations is very thorough,
gneissic rock mass at the site was generally saturated to this assumption is of course suspect and a more com-
within a few metres of the surface, and Lugeon-type plicated picture may arise. When reservoir impounding
permeability tests in the four investigation boreholes begins there are likely to be significant reductions in
indicated values in the range 1 to 0.01 Lugeons, mostly effective stress near the dam, and fluctuations of the
0.1. An inversion of the median Lugeon value is close latter, usually on a seasonal basis.
to the mean Q-value, the significance of which will be Savich et al., 1983 noted the above effects when the
clarified in Chapter 9. 270 m high Inguri arch dam was being constructed. A
Relationships between rock quality, depth and seismic velocity 89

(a)

(a)

(b)

Figure 5.30 a) Hydraulic fracturing based stress estimation at


Gjøvik. b) Velocity depth trend next to one of the
boreholes where seismic tomography was performed.
RQD and Fm1 (and also Q-value) did not show
improvement with depth. Kristiansen, 1991 pers.
comm., Barton et al., 1994.

(b)

Figure 5.31 Vp distribution in Inguri dam abutment, and records


sharp change in deformation was noted when impound-
of Vp changes with construction, reservoir filling and
ing of water began. Significantly, Vp values reduced (Vs
seasonal fluctuations. Savich et al., 1987.
values also reduced), and there was a reduction in resis-
tivity, each implying reduced effective stresses due to
the rise of joint water pressures in spite of the grouting four years of performance monitoring, since their last
and drainage measures. The most intensive reductions referred article (Savitch et al., 1983). Figure 5.31a shows
in Vp occurred during periods of water level rise, with the deeply penetrating lower velocity contours in the left
weaker reductions, when the water level was constant. abutment and a joint pole concentration plot. Figure
Interestingly, the ‘conventional’ effective stress type link- 5.31b shows the complex coupling of Vp and reservoir
age with Vp was modified, and radically changed in sub- level response, with predominant gradients showing Vp
sequent long term monitoring. increasing with reservoir level (H).
Savitch et al., 1987 gave updated and longer term A nine-year record of Vp, percentage change of Vp, dam
syntheses of the Inguri dam response, following another load and reservoir level fluctuation shown in Figure 5.32b
90 Rock quality, seismic velocity, attenuation and anisotropy

confirms the general, but slightly delayed increase in Vp


with reservoir level, following the initial reduction in Vp
which appeared to be effective stress related. A more
detailed look at Vp–H response, with smoothed-out
behavioural trends (Figure 5.32b) shows a small delay.
Savitch et al., 1987 interpret the near-synchronous
behaviour (after first reservoir filling) as closure of
joints or cracks due to the reservoir load on the dam.
However, at higher pressures perhaps a widening of
large cracks occurs. Reservoir draw-down rate appears
to be important since if the draw-down rate exceeds the
‘permeability velocity’, a sharp decrease in Vp occurs,
corresponding to the classic effective stress model.
Before leaving this chapter on effective stress-depth
(a)
effects, we will review an interesting set of cases presented
by Moos and Zoback, 1983. Seismic velocity measure-
ments in four wells to roughly 1 km depth, can serve as a
bridge to the much deeper seismic velocity-depth profiles
that are given in Chapter 11.
Four wells, varying in depth from 0.6 to 1.2 km were
surveyed with borehole televiewer, and sonic logged,
which involved measuring vertical travel times over a
1.2 to 1.5 m interval. One of the wells (the Crystallaire
well, termed XTLR, was drilled 4 km from the San (b)
Andreas fault, in crystalline rocks, in the Mojave Desert
area. This well was the subject of three separate seismic Figure 5.32 Ten years of recording P-wave velocity and reservoir
fluctuations at Inguri dam. Top) Ultrasonic, borehole
investigations using sonic logging (from below the
P-wave velocity fluctuations, ‘relative velocities’, and
water table), vertical seismic profiling (VSP) using an
reservoir levels, including filling. Bottom) Seismic P-
air gun source, and nearby seismic refraction. wave velocity fluctuations show slight ‘inertial delay’ rel-
The three different frequencies were respectively ative to the seasonal reservoir fluctuations. Savich et al.,
2 MHz, 75 Hz and 10 Hz, and these meant three 1987.
fundamentally different wave lengths in relation to the
fractures. Wave lengths were respectively 20 to 60 cm
(i.e. similar to the fracture spacing), 60 m (using a 30 m
geophone interval) and several hundreds of metres. In Another well reported by Moos and Zoback, 1983,
addition, laboratory samples were taken at intervals showed reduced depth-dependent velocities in relation
down the core, and were subjected to appropriate con- to the above, due to the somewhat lower fracture
fining pressures, based on an assumed effective stress frequency. This was the MONT-1well, from the
gradient of 16.7 MPa/km, designed to correspond to Monticello reservoir in South Carolina, USA, drilled
the effective overburden stress. These four sets of data through granodiorites. Joint or fracture frequencies and
are compared in Figure 5.33. The majority of the frac- P-wave velocities are shown together with Vs and Vp/Vs
tures (or joint sets) were steeply dipping. ratios in Figure 5.35. The solid circles representing the
Overall there was remarkable agreement between the ultrasonic tests on the laboratory samples were in this case
three in situ methods, but the ultrasonic measurements very close to the in situ sonic-log data for this sparsely
on the intact cores gave significantly higher velocities; fractured well.
about 2 km/s higher at shallow depth and about 1 km/s These two contrasting wells showed increases of aver-
higher at intermediate depths. There was a relatively age Vp from about 3.5 to 5.2 km/s (XTLR, Fm1  1
high degree of fracturing in this well, as shown in to 4) and 5.4 to 6.1 km/s (MONT-1, Fm1  1 to 2),
Figure 5.34, which also shows the detailed sonic log, over the respective 100 to 850 m and 50 to 1100 m
and its ‘divergency’ from the lab (intact) data, as the depth ranges. Naturally, almost every fracture zone (or
surface is approached. joint swarm) was a zone of low velocity.
Relationships between rock quality, depth and seismic velocity 91

Figure 5.34 Details of joint frequencies in the XTLR well, show-


ing generally small variations with depth except in
Figure 5.33 Comparison of velocity-depth trends obtained from upper and lower 50 m, together with the generally
three different frequencies of field measurement increasing trend of the sonic log velocities with depth.
(smoothed sonic log, VSP, refraction seismic) with Moos and Zoback, 1983.
ultrasonic tests on intact laboratory samples, tested at
appropriately increasing effective confining stress
levels. XTLR well in crystalline rocks, 4 km from San The average velocity-depth gradients shown in
Andreas fault zone. Moos and Zoback, 1983. Figures 5.33 (XTLR well) and in Figure 5.35 (MONT-1
well) are as follows:

Besides the description of joint frequency, there was  Vp  5.2  3.5


relatively little information in Moos and Zoback’s inter- XTLR  
   2.3 s1
esting article from which to judge the number of joint  depth  0.75
mean
sets or the joint character. However, an attempt can be
made to match these quite different velocity-depth
gradients with a deconvoluted version of Figure 5.36.  Vp  6.1  5.4
This is shown in Figure 5.37, and will be explained MONT  1     0.7 s1
 depth  1.05
shortly. mean
92 Rock quality, seismic velocity, attenuation and anisotropy

Two other wells reported by Moos and Zoback, 1983,


showed little, or in one case reversed (negative Vp/depth
gradient) due to the increasing frequency of fractures
from about 350 to 600 m in the well. Velocities in the top
200 m averaged about 5.5 km/s. When little Vp-depth
sensitivity is shown, it suggests few fractures. In fact the
range of average Vp for a well named MONT-2 was as
high as 5.9 to 6.1 km/s despite medium fracture density.
This suggests (in Q-value terms) that there might have
been only one set of joints, giving a high Q-value, and
further that the joints and rock had characteristics allow-
ing for stiff, possibly smooth-walled apertures.

5.5 Integration of velocity, rock


mass quality, porosity, stress,
strength, deformability

It has been shown in previous sections how the P-wave


velocity is sensitive to each of the factors listed in the
above sub-title. To this we must also add moisture con-
Figure 5.35 Velocity-depth trends were less marked in the tent (for the matrix) and ground water level (for the rock
MONT-1 well, in less jointed granodiorites, where mass). The assumption will be made in the following
laboratory sample velocities were also close to the development that seismic velocity measurements will
field result. Moos and Zoback, 1983. most frequently be made in saturated rock masses. The

Figure 5.36 Integration of rock quality Q-Vp-Emass in a model that incorporates depth, porosity and rock strength adjustments. Note that
Emass (or M) represents the static modulus of deformation, from plate loading tests and from back-analysis of measured defor-
mations. Vp is the seismic velocity measured from refraction seismic, and from cross-hole seismic tomography, in the case of
greater depths. Barton, 1995.
Relationships between rock quality, depth and seismic velocity 93

correlations developed will be based on this assumption, This correction is necessary because the rock quality
and systematic errors will of course arise if drainage Q-value was originally developed in 1974, for correl-
causes drying out of the matrix and/or joint water. ation with tunnel and cavern rock reinforcement and
Since velocity–rock quality correlation is a complex support needs (i.e. rock bolts and sprayed concrete,
task, no hesitation must be made in adding some degrees respectively). The rock quality Q-value only uses the
of sophistication to the simple model Vp  3.5  ratio strength/stress ( c/ 1 in the SRF factor – see
log10Q proposed earlier (Figure 5.10). Avoidance of Appendix A), when major principal stress levels (and
mathematics suggests the use of a graphic method for their concentration as maximum tangential stress 
converting the formulation for hard, unweathered, low- 3 1 – 3) are causing stress-related fracturing.
porosity, near-surface rock masses (i.e. typical Sjøgren It is probable that in a tunnel excavation disturbed
et al., 1979, data) to conditions towards the other end of zone (EDZ), the potentially large values of SRF (that
the seismic and rock quality scale, e.g., low strength, reduce the Q-value directly) can also be used in princi-
weathered, high porosity, highly stressed (or unstressed) pal to predict the measured reductions in velocity and
rock masses. deformation modulus that are typically recorded in the
The development shown in Figure 5.36 which was walls of deep shafts and tunnels (e.g. Barton and
introduced by Barton 1995 and 1996a, has opposing Bakhtar, 1983 who back-calculated moduli of 3.5 GPa
corrections for porosity and depth (i.e. stress) since in the outer 3 m, or one radius, of a 1,600 m deep shaft
these cause opposing influences on velocity. In addi- in steeply bedded, highly stressed, jointed quartzites,
tion, an adjustment for uniaxial compression strengths compared to 65 GPa at a depth of two shaft diame-
different from a typical hard rock 100 MPa (or more) is ters, using multiple position borehole extensometers, or
made by the following simple normalisation of the rock MPBX).
quality Q-value: In highly stressed cases, c will tend to be high for
hard massive rocks subject to dynamic and sometimes
explosive stress-slabbing, and c will be low for soft
c (5.6) rocks that are subject to a slower, ‘plastic-deformation’
Qc  Q  ( c expressed in MPa)
100 type of squeezing. In equation 5.6, c/100 corrects the

Figure 5.37 Conversion of the Barton, 1995 rock quality Qc-Vp model of Fig. 5.36, into a more familiar velocity-depth format. Note that
there will be a tendency for ‘curve-jumping’ (i.e. ‘Q-jumping’), as a near-surface rock quality improves at greater depth. This
will be due to the reduced effects of weathering, and due to a tendency for reduced jointing frequency. Note the contrasting
directions of the ‘N’ and ‘J’ arrows shown in the figure, together with the s1 (km/s/km or m/s/m) gradients.
94 Rock quality, seismic velocity, attenuation and anisotropy

Q-value to an approximately suitable value of Q c to (Figure 5.36, right-hand column) become closer to seis-
correlate more closely with the rock mass static deform- mically derived dynamic E moduli (from Vp, Vs and
ation modulus (M, or Emass) and with the seismic veloc- density, see Chapter 1). If truly undisturbed static mod-
ity Vp, particularly for the case of softer rocks. ulus testing could be achieved, the normal discrepancy
The ‘opposed corrections’ given by the two sets of between static deformation modulus and dynamic E
diagonal lines in Figure 5.36 are designed to do the modulus might be lessened, despite the fundamentally
following: different levels of strain involved in each case.
An illustration of application of the Q–Vp–M
1. A strongly non-linear initial correction for depth
method to very soft rocks can be given here, based on
gives greater sensitivity to ‘acoustic joint closure’
Q-logging of tunnels in chalk marl (Terlingham Tunnel,
for weaker and lower quality rock masses.
Beaumont Tunnel, and UK sector Channel Tunnels,
2. A weakly non-linear porosity correction also gives
and selected marine drill core: PB1 to PB8). Details of
larger changes of velocity for the weakest rocks.
the rock quality Q-logging are given by Barton and
The chart, which should be considered as an approxi- Warren, 1996c.
mate engineering guideline, was developed with the A weighted mean value of Q  8 was obtained from
hard rock (Vp  3.5  log10Q) relationship (Figure the so-called ‘precedent study’ of nearby tunnels in
5.10) as a ‘core’ (see diagonally-aligned black discs in chalk marl, and from marine drillcore. This Q-value
Figure 5.36). Development for soft rocks occurred by a was found to compare closely with the overall mean
process of trial-and-error fitting of Q, Vp, c, n and Q-value recorded in the contractor-owner (Trans
depth data from known sites in jointed chalks, (Saul Manche Link – EuroTunnel) TBM face logs from the
Denekamp, personal communication), jointed chalk 20–24 km subsea chainage near the English south
marl from the UK end of the Channel Tunnel, sand- coast, where an overall mean value of Q  9 was
stones, mudstones, shales, welded tuff and ignimbrite. obtained from the three machine bored tunnels. In this
Data from Israel, England, Japan and China were zone, significant tunnelling problems were caused by
included. Depths ranged from about 25 meters to more the (salt) water leakage and overbreak, hindering
than 1000 metres in the case of tuff/ignimbrite from PC-element ring-building, and causing problems with
UK Nirex’s Sellafield site, where cross-hole tomography electrical equipment on the TBM.
and laboratory tests could be compared with NGI’s The mean c value for the chalk marl was 6 MPa.
detailed rock quality Q-logging (and joint index test- Thus from equation 5.6 we have:
ing) of 9 km of drill core (see Chapter 4).
The first two empirical relationships listed in the top
6
of Figure 5.36 were derived from extensive field test data Q  8, c  6 MPa, Q c  8   0.48
for hard rocks (Barton, 1996). Testing with soft rocks has 100
shown that the modified Qc term gives satisfactory fit,
which is improved when the porosity and depth correc- This Qc value intersects the central diagonal line (equa-
tions are also made. Thus we have the following approx- tion 5.7) in Figure 5.36 at Vp  3.2 km/s. Correction for
imations for general use. Note that both Vp and average porosity (n  27.7%) results in a reduction of
deformation modulus (M or Emass) are predicted to 1.6 km/s giving 3.2–1.6  1.6 km/s. Tunnel depths of,
increase with depth (see detailed discussion in Barton, for example 40 m, bring this value up to about 2.0 or
2002a). 2.1 km/s. Offshore geophysics carried out during several
campaigns indicated P-wave velocities generally in the
Vp  log10 Q c  3.5 (km/s) (5.7) range 2.0 to 2.6 km/s for the UK chalk marl. A Q-value
of about 20 is needed to explain the upper velocity of
2.6 km/s using the above method. This is in line with the
M(mean)  10Q c1 3 (GPa) (5.8) otherwise generally good rock mass Q-values, registered
outside the 4 km zone with much overbreak.
The predicted moduli of M  0.5 GPa (minimum)
Although there is little data for deformation modulus and 3 GPa (mean) compare with (disturbed) but
measurement at ‘undisturbed’ depths of hundreds of unjointed laboratory scale moduli of 0.64 GPa (mean) and
metres, it will be noted that the predicted static moduli a range of 0.15 to 4.2 GPa. Deformation measurements
Relationships between rock quality, depth and seismic velocity 95

in the tunnels were interpreted as indicating an million cubic meter post-glacial landslide, at Koefels in
in situ modulus of about 1 to 2.5 GPa for a range of Austria, which covers a valley over an area of some
rock qualities, quite consistent with the above. 10 km2. Brueckl and Parotidis, 2001, found that the
The correction Q c  Q  c/100 can also be landslide mass could be approximated by a Vp-depth
applied in the direction of very hard rock, to adjust the model which we can simplify to Vp  400D0.32 m/s. So
value of Q c above that of the logged Q-value. As we at 50 m, 100 m, 200 m and 300 m (the average depth),
shall see in Chapter 7 on EDZ phenomena, this gives a the velocities would, in round figures be approximately
useful hard rock correlation between measured veloci- as low as 1,400, 1,750, 2,300 and 2,500 m/s, compared
ties and observed Q-values from greater depths. We will to a bedrock of 5.2 km/s.
examine predicted and measured velocities and moduli As a first approximation one might view the Qc -Vp –
from greater depths in Chapter 7. depth model of Figure 5.37 and consider a faulted-
The sloping lines for depth adjustments that are crushed-rock Q-value of 0.01, which would reduce to a
shown in Figure 5.36 can be converted to velocity-depth Q c – estimate of 0.001 if weakened rock of say, 10 MPa
gradients, with a format more familiar to geophysicists, uniaxial strength was involved. The relevant Vp – depth
as shown in Figure 5.37. Here one may note the marked curve gives a certain geometric similarity to the above
linearity at depths beyond the first 250 m, and the velocity increases down to 300 m depth, but needs a
marked non-linearity in the upper 100 m in particular. minor ‘parallel shift’ to somewhat lower values.
With awareness of the different velocity scales, we Such could be ‘achieved’ by a) a minor downwards
can compare the velocity-depth trends of the XTLR adjustment of velocity, due to a porosity correction
well near the San Andreas fault (Figure 5.34) with the (Vp) for slightly altered rock, or b) by a larger down-
overall gradients of a rock mass quality of Qc  0.1 in wards adjustment due to the larger mass porosity of the
Figure 5.37. This suggests a well tectonised rock mass failed materials, e.g. 20%, but then necessarily from a
with several joint sets, and possibly with clay coatings higher Q c value as a starting point. Detective work,
and smooth or slickensided joints. Unfortunately, the preferably aided by local knowledge of the site, may
detailed condition of core seems seldom to be described result in improved insight, where a hint concerning the
in published articles, so whether this ‘picture’ is realistic deformation modulus of the slide masses (about 5 GPa)
is uncertain. might also need to be used.
At the extreme end of the rock mass quality spec-
trum, one may refer to the seismic studies of a several
6 Deformation moduli and
seismic velocities

Although it was not the original intention in this book to the foundation using larger scale cross-hole, between-
explore in detail the ‘static’ deformability of rock masses, gallery, or VSP (hole to surface or tunnel perimeter)
a superficial treatment is necessary, in view of the com- velocity measurements. Classic examples of this were
mon use of seismic measurements to extrapolate near- shown in Figure 4.1 and 4.2.
surface deformation or elastic modulus measurements. A useful basic comparison between the ‘static’ modu-
We saw in the last chapter how the loading of individ- lus of deformation Ed (or M in Figure 5.36) and the
ual joints, or of multiple jointed columns of rock in the so-called modulus of elasticity (Ee) is shown in Figure 6.1.
laboratory created changes in the P-wave and S-wave The modulus of elasticity (Ee) is traditionally obtained
velocities and amplitudes. Detailed monitoring of from the gradient of the unloading curves, which are
in situ loading tests using sonic logging in boreholes often similar, and supposedly have elastic character due
beneath plate loading tests, such as shown in Chapter 3 to the frequent closed state of the stress-deformation
(Figure 3.15), also shows correlation with the moduli loops. These results were given in the 1st ISRM Congress
obtained. However, due to the inevitable damage in the by Kujundzíc and Grujíc, 1966. Three of their figures
unloaded zone around the sites of the tests, the lowest are reproduced in Figure 6.1.
moduli (and lowest velocities) are usually registered clos- The total deformation measured at the highest load
est to the loading plates, flatjacks or pressure chambers, level, after several loading cycles, is the usual basis for the
while higher velocities and moduli are registered at greater calculation of Ed (Figure 6.1a). Thus defined, Ed may
depth. This is probably due to the more uniform stresses change with stress level, while Ee is considered a con-
outside the excavation disturbed zone, or EDZ, where stant. Partly for this reason, the inequality Ee  Ed can
shear stresses are less due to the radial stress ( r) being be quite large, as it depends also on the stress level, which
higher. There is increasing discrepancy between dynamic in turn is usually based on the size and type of dam (or
and ‘static’ deformation moduli in the near-surface, low other structure) to be located on the particular rock
quality, weathered zone, that may unfortunately need foundation.
to be used as a dam or large building foundation. The basic inequality of Ed and Ee is clearly shown in
In this chapter there will be detailed treatment and Figure 6.1b and c. Kujundzíc and Grujíc, 1966, assem-
examples of the multitude of ‘moduli’ that have prolif- bled most of their data from tests on limestone founda-
erated with the combined use of loading (and unload- tions at Yugoslavian dam sites carried out in the 1950s
ing) ‘static’ moduli, and the attempts to extrapolate or and 1960s. As can be seen from the data, the largest
compare these with the ‘dynamic’ moduli, both from inequalities (ratios of 1.5 to 2.5) are seen at the lowest
laboratory samples and in situ measurement. moduli values (5 GPa). The values of Ee and Ed
followed approximately the following trend:

6.1 Correlating Vp with the ‘static’ 15.50 Ed


moduli from deformation tests Ee  (6.1)
6.45  Ed
The method that is perhaps most common for relating
seismic velocity and in situ modulus measurements, is where Ee and Ed are expressed in GPa. Let us look at two
what we may call the seismic characterisation method. The examples to illustrate the trends of Ee  Ed, with dif-
rock mass at, or beneath, the site of a deformability test ferences reducing when higher quality rock masses are
is logged by downhole, cross-hole or shallow refraction involved. Assuming Ed of 5 and 50 GPa, we see from
seismic. On the basis of the local Vp value the measured equation 6.1 that the elastic moduli are predicted to be
moduli are extrapolated (or interpolated) to other parts of 8.9 and 57.3 GPa.
98 Rock quality, seismic velocity, attenuation and anisotropy

Note that these larger differences with Ee  Ed at lower


qualities, are largely caused by an increasing hysteresis
in the load-unload behaviour of the individual rock
joints, as rock mass quality deteriorates, typically due to
weathering and unloading effects near the surface. In
Chapter 16 we will also see the different shapes of these
rock mass loading curves caused by predominance of
joint closure (concave type) over shear (convex type). A
combination of both components may give a semblance
of ‘linearity’. (Barton, 1986).
(The above inequality Ee  Ed is in the ‘static’ loading,
ultra-low-frequency sphere of geophysics. This behaviour
is not directly related to another key inequality in geo-
physics, namely that the (static) joint stiffnesses become less
than the inverse of the (dynamic) joint compliances, as
rock mass quality deteriorates. These aspects are discussed
in Chapters 15 and 16 in relation to shear-wave splitting).
(a) Grujíc, 1974, also reported results of Vp measurements
in the galleries at a Yugoslavian dam site, as shown in
Figure 6.2. The ability to use Vp measurements to extrapo-
late flatjack or plate loading test results is obviously
an important part of the economy and thoroughness
of major foundation studies. In most cases, working
hypotheses are established that relate deformation mod-
uli with the seismic velocities. In the case of the Mratinje
dam on limestone foundations, Grujíc found the fol-
lowing relationship:

Ed  9.1  4.8 (Vp  3.6) GPa (6.2)

With Vp  4.5 km/s, (an implied nominal 25 m depth


rock quality Qc  10), this equation suggests a static
deformation modulus Ed of 13.4 GPa. This is midway
between the minimum and mean M-values shown in
(b) Figure 5.36. The tendency for ‘disturbed zone’, near-
valley-side reductions in Vp and presumed reductions in
Ed (or M) can be seen in Figure 6.2.
The basis for the above correlation between Vp and
Ed, were the so-called polar measurements of Vp sur-
rounding the flatjack (and therefore loaded) locations.
These are illustrated in plan and in vertical section, in
Figure 6.3. The upper line in Figure 6.3c shows Ee (E),
and the lower line shows Ed (D), and their correl-
ations with velocity. These obviously relate to velocities
of the flatjack-loaded rock mass, giving a range of Vp,
together with rock quality variations, from 1.5 to
(c)
5.5 km/s. The rocks involved were variously jointed
Figure 6.1 Definition of Ed and Ee, and examples of their differ- and weathered Triassic limestones.
ences, from limestone dam foundations in Yugoslavia. A very thorough review of inter-relationships between
Kujundzíc and Grujíc, 1966. seismic velocity measurements and E-moduli and
Deformation moduli and seismic velocities 99

Figure 6.2 Example of Vp measurements in exploration galleries at a dam site in Yugoslavia. These are used to economically extrapolate
deformation modulus test results to other parts of the dam foundation. Grujíc, 1974.

(a)

(b) (c)

Figure 6.3 Flatjack deformation tests for determining the moduli D(or Ed), and E (or Ee), and correlation with velocity in the Triassic lime-
stones. Grujíc, 1974.

D-moduli obtained from plate loading tests in Italy was anisotropy, which the author nevertheless concluded,
given by Ribacchi, 1988. These are shown in Figure 6.4. was greatest with samples that had been recovered for
The particular emphasis of this review article was laboratory tests. The results of in situ investigations at
rock mass deformability, and complexities caused by eleven sites were given. These involved several varieties
100 Rock quality, seismic velocity, attenuation and anisotropy

Ee GPa Ee GPa

(a) (c)

Ed GPa
Ed GPa

(b) (d)

Figure 6.4 a,b) Results of in situ deformation tests at eleven sites, showing Ee and Ed correlated with Vp2. c,d) Mean data from each site shows
reduced scatter. Note dynamic Poisson’s ratio gradients Ribacchi, 1988.

of limestones, schists, gneisses, granodiorites, mylonites, conducted perpendicular and parallel to layering, bed-
dolomite, a high porosity (n  30%) calcarenite, and ding, or schistocity, as appropriate.
inter-bedded sandstones and mudstones. Both the plate The author cited examples of sites that showed
loading tests and the seismic velocity measurements were good correlation between modulus and typical indices
Deformation moduli and seismic velocities 101

of jointing (RQD, and joint frequency Fm1). The scat-


ter of more than 100 data points for modulus E or D
versus V2p in Figure 6.4 is quite large for the whole sample,
but reportedly showed good correlation for limestones
and dolomites. Figures 6.4a and b show all data points
for Ee and Ed, while somewhat reduced scatter for the
mean data at each site is shown in Figures 6.4c and d.
Ribacchi’s data from a limestone and dolomite site are
shown in Figures 6.5a and b. Here, modulus is plotted
on a log scale, and velocity is not squared as above,
thereby showing much narrower scatter. Trend lines for
these and other limestones (Yugoslavian) are shown in
Figure 6.6.
The deformation moduli M(mean) and M(min.)
tabulated in the inset to Figure 5.36 (in Chapter 5) are
also encompassed by the above trend lines. The log-
linear equation given in Figure 5.36 simplifies to:

M  10(Vp  0.5)/3 (GPa ) (6.3)


(a)
and serves as a mean value for the in situ E-moduli (Ee
shown in Figure 6.5a), while it is closer to an upper-
bound for the in situ deformation moduli (Ed shown in
Figure 6.5b). The M(min.) values given in Figure 5.36
independently encompass the minimum Ed trend lines
given in Ribbacchi’s Figure 6.6.
Kikuchi et al., 1982, also presented a very compre-
hensive set of modulus and velocity data, this time from
investigations at Japanese dam sites. They established a
rock grading system based on a combination of seismic
refraction data, deformation moduli from plate load
tests, Mohr Coulomb parameters from in situ shear
tests, and rock rebound hammer (modified to 30 mm
impact diameter for soft rocks). Their data extended
from weathered or soft rocks with uniaxial strengths as
low as 2 MPa up to extremely hard rocks with uniaxial
strengths as high as 300 MPa. The in situ P-wave veloci-
ties ranged from 0.4 up to 5 km/s.
Figure 6.7a and b show examples of their Vp–rock
grade, and Vp-D (modulus of deformation) correlations,
while Figure 6.7c gives the well-correlated log E–Vp rela-
tionship. The elastic (E) modulus and the deformation (b)
(D) modulus were defined as tangent and secant gradi- Figure 6.5 Vp versus logarithm of deformation moduli Ee and Ed, for
ents at the maximum loading of 7 MPa. Their best-fit two Italian dam sites in limestone and dolomite. Ribacchi,
trend lines for E and D compare quite closely with 1988.
envelopes 2 or 3 and 1, respectively in Figure 6.6 (from
Ribacchi, 1988). two similar sites consisting of crystalline limestones and
Broad correlations between seismic P-wave velocities marly limestones with innumerable clayey interlayers of
and in situ loading test results at dam sites are also millimetre thickness produced the results shown in
demonstrated by Navalón et al., 1987. Investigations at Table 6.1 (M  deformation modulus).
102 Rock quality, seismic velocity, attenuation and anisotropy

These three depth zones were defined at each of


the two dam sites, and were relevant to foundation
testing in the left and right abutments and in the
riverbed at each site. Similar geology was involved in
each case.
The above low velocity data broadly fits the simple
empirical equation developed in Chapter 5 (Figure 5.36):

D or M  10Q
1
3 GPa (6.4)

when Vp values are converted to Q-values using the


‘hard rock’ relation Vp  3.5  log10Q. However, the
‘deep zone’ data with Vp  3.5 km/s have surprisingly
high moduli (approx. 14 to 20 GPa) which are higher
than that predictable from equation 6.4. For some rea-
son the moduli but not the velocities appear to be
(a) increased by the ‘deep zone’ boundary condition (i.e. by
higher stress).
The process of rock mass loading and even failure
under various sizes of plate loading (0.28 to 0.45 m in
diameter) was monitored by Savitch et al., 1974, using
simultaneous Vp measurement in the first one metre
beneath the plates. The testing was performed at the
Inguri arch dam in Russia, which is founded on medium
and intensely jointed limestones. The range of deform-
ation moduli was 1 to 15.7 GPa.
Increments of loading were applied, to final levels of
about 10, 20, 30 or 50 MPa (depending on rock qual-
ity) which caused characteristic rises and sharp falls of
Vp-values (Figure 6.8a, b, c) for various rock qualities,
which were sampled by arranging the loading tests at
different depths below the surface. The Vp monitoring
was performed cross-hole, to 1 m depth below the plates,
in boreholes that were located close to the edges of the
plates, across a diameter.
(b) Figure 6.8 shows small arrows on each set of curves.
Figure 6.6 Vp versus logarithm of deformation moduli Ee and Ed, Those arrows pointing downwards signify ‘failure’ as
from Italian, Yugoslavian and American in situ test evaluated by the seismic results, while the arrows below
data. See Ribacchi, 1988 for references. the curves signify failure as assessed by the accelerating

Table 6.1 Deformation moduli and velocities from numerous plate load tests. Navalón et al., 1987.

A) Vp (km/s) M (GPa) B) Vp (km/s) M (GPa) C) Vp (km/s) M (GPa)

Decompression zones 0.7–2.0 0.4–3.9 Intermediate zones 1.8–2.2 3.4–4.9 Deep zones 3.3–3.6 17.6–19.7
0.9–2.3 0.7–5.9 2.6–2.7 7.4–8.3 3.5 14.2–16.2
1.0–2.4 0.9–6.4 2.6 7.3 3.5 16.2
1.5–2.5 2.0–6.4 2.6 7.4
Deformation moduli and seismic velocities 103

(b)

(a)

(c)

Figure 6.7 Correlations of refraction seismic velocities with weathering grade, deformation modulus and elastic modulus from Japanese dam
sites. Kikuchi et al., 1982.

deformation. Consistent relationships between deform- with stress increase (uniaxial or triaxial), but Shea and
ation modulus, velocity and the ultimate strengths of Hanson, 1988, identified two other phases of behaviour
the rock masses are shown by the following tabulation: as well which resemble those we have just seen in
Savitch et al., 1974 plate load tests.
Table 6.2 Plate load tests taken to ultimate failure. Savitch et al., Phase I represented the rapid rise in P-wave velocities
1974. due to ‘closing of layer cavities’ at quite low levels of load-
ing (MPa assumed). Attenuation also decreased in this
Test No. D (GPa) Vp (km/s) c (mass) MPa
phase as shown in Figure 6.9a Phase III marked the
3 0.9 1.3–2.2 2–3 decline of these two trends, while Phase II represented the
6 1.0 2.2–2.4 3–5 increase in attenuation that probably signified the creation
4 2.8 3.6–3.9 19–24 of micro-cracks. There was hardly any increase in P-wave
1 3.4 3.6–3.8 22–23 velocity in this phase, although the S-wave velocity con-
7 15.4 3.9–4.2 42–78 tinued to increase.
Figure 6.9b shows that triaxial conditions (1.7 MPa
confinement) caused consistently reducing attenuation
Coal, which in some ways resembles jointed rock at in Phase I and II, then a sudden increase in attenu-
reduced scale, also shows the expected increases in velocity ation and reduction in velocity as failure approached in
104 Rock quality, seismic velocity, attenuation and anisotropy

Figure 6.8 Effect of stress on Vp values beneath plate load tests at the Inguri arch dam in Russia. Savich et al., 1974.

elastic moduli to rock engineering design will also be


addressed. The four standard equations, two of which
were introduced earlier in Chapter 1 are reproduced
here together, for ready reference (ISRM, 1998).
Dynamic Young’s modulus Edyn:

3( VP / Vs )2  4 (6.5)
E dyn  Vs2
( VP / Vs )2  1

(1  d )(1  2d ) (6.6)


 Vp2
(a) 1  d

Dynamic Poisson’s ratio d:

1 ( VP / Vs )2  2 (6.7)
d 
2 ( VP / Vs )2  1
(b)

Figure 6.9 a) Uniaxial, and b) triaxial test of coal, showing velocity


and attenuation changes caused by cleat and microcrack Dynamic shear modulus
:
behaviour. Shea and Hansen, 1988. (also with symbol G)

Phase III. Broadly similar (but inverted) behaviour of E dyn


the P-waves was also seen.
 Vs2  (6.8)
2  2 d

6.2 Dynamic moduli and their Dynamic bulk modulus Kbulk:


relationship to static moduli

In this section, which principally concerns dynamic K bulk  ( Vp2  4/3 Vs2 ) (6.9)
moduli and dynamic Poisson’s ratio, the use of Vp and
Vs measurements to derive the four standard dynamic
E dyn
elastic properties of rock masses will be discussed. The  (6.10)
validity or otherwise of these dynamic, small-strain, 3  6d
Deformation moduli and seismic velocities 105

Figure 6.11 Ratio of Vp and Vs at a hard igneous rock site in


Norway. Vp/Vs  1.8–1.9. The full range of dynamic
Poisson’s ratios was from 0.15 to 0.39, with higher
Figure 6.10 Contrasting laboratory and field data for the Vp and values when velocity was lower. Sjøgren et al., 1979.
Vs values of limestones, with calculated dynamic
Poisson’s ratios. Note the over-riding tendency of
higher dynamic Poisson’s ratio in the case of the lower cases, values in excess of 1.0 have been registered. This
velocity in situ data. Ribacchi, 1988. occurs as shear failure is approached, in the case of biax-
ially loaded model rock masses, having two conjugate
fracture sets that are under significant levels of shear
It will be noted that if the dynamic Poisson’s ratio is stress. (Barton and Hansteen, 1979, Barton, 1993a,
estimated (rather than derived from Vp and Vs), the three and Barton, 2004b).
dynamic moduli can theoretically be estimated from Vp Measurements of Vp and Vs at an unweathered site in
measurement alone. As pointed out later, this can cause Norway are shown in Figure 6.11, and indicate ratios of
significant inaccuracies, and such values given in the Vp/Vs of about 1.8 to 1.9. The corresponding dynamic
literature should be treated with caution. Poisson’s ratios were found to lie in a narrow range, more
The manner in which Vs and Vp values are distrib- than 80% of the values were between 0.26 and 0.32. The
uted in relation to the general quality of the rock mass mean value of 0.28 in fact lay close to the maximum
was illustrated in Chapter 1 (Figure 1.7 from Sjøgren, RQD and minimum joint frequency trend for this par-
1984). Further sets of data, both from laboratory sam- ticular site, and corresponded to a ratio Vp/Vs between
ples and from comparable field data (Ribacchi, 1988) 1.8 and 1.85. In the rock masses investigated, the
are illustrated in Figure 6.10. authors found that the full range of dynamic Poisson’s
Ribacchi’s data shows a particularly clear demarcation ratios was from 0.15 to 0.39.
between the lower dynamic Poisson’s ratio in the case of Deere et al., 1967, addressed the important differences
the higher velocity laboratory data, and the opposite between field measurements of EF dyn and laboratory
trend for the lower velocity field data. High dynamic measurements of EL dyn of the intact rock, by utilising
Poisson’s ratio are a clear sign of the influence of joint- their observation that (VField/VLab)2 resembled RQD/
ing. In shear zones and faulted rock, high values of d 100. (This was referred to in Chapter 5, Table 5.1.) This
are common. was based on the Onodera, 1963, suggestion of using the
During ‘static’ flatjack biaxial loading tests of rock field/lab velocity ratio VF/VL as a measure of rock quality.
masses, ‘static Poisson’s ratios’ (or lateral expansion coef- The resulting method for estimating EL dyn/EF dyn is
ficients) in excess of 0.5 may even be measured. In special shown in Figure 6.12. As can be noted from the spread of
106 Rock quality, seismic velocity, attenuation and anisotropy

data, the inverted knee-shaped trend does not give a good


method for estimating the lowest values of the ratio, as
RQD and (VF/VL)2 vary too much in the region EL dyn/
EF dyn  0.1–0.3. In higher quality rock masses however,
a more consistent trend is observed, some of the data
from the USA even indicating EF dyn  EL dyn , for values
of (VF/VL)2 as high as 0.9, or RQD  90% or more.
(a) When commenting on the differences between static
and dynamic moduli, Wang and Nur, 1992, make the
usual observation concerning the different strain ampli-
tudes involved (perhaps 103 and 106 respectively)
but also point out that the different moduli are thereby
directly caused by the presence of pores, cracks (and
joints). These are deformed in static tests, but hardly
deformed by dynamic waves, but when stresses are very
high and pores, cracks and joints are (almost) closed, the
static and dynamic moduli are likely to be very close.
The possibility of estimating Edyn values from Vp meas-
urement alone, by just estimating the value of the
dynamic Poisson’s ratio (d) instead of measuring both
Vp and Vs was referred to earlier. This would involve
using equation 6.6 instead of the correct method utilising
equation 6.5. In warning against this short-cut, Stacey,
1977, assembled what he called ‘reliable’ Edyn values from
the literature. These values are shown in Figure 6.13a.
(For the numbered references see Stacey, 1977).
(b) Stacey also assembled a large number of Ee-Edyn data
(Figure 6.13b) and Ed-Edyn data (Figure 6.13c). These
two figures show the apparently irrelevant nature of Edyn
in comparison to the standard rock engineering methods
of testing deformation modulus for design purposes.
The doubt nevertheless remains that most of our large
scale methods are testing an excavation disturbed zone, or
EDZ, rather than the undisturbed state, with its higher
and maybe more isotropic in situ stresses.
There is a contrary factor that most rock masses
observed or tested, are actually on a major unloading
curve due to erosion or due to rock excavation, and may
have correspondingly higher joint stiffnesses, than if
loaded up without this prior unloading.
The mismatch of the static and dynamic moduli in
jointed rock masses, except where rock qualities are very
high and strains are very small, is probably a universal
rule, unless extremely small strains are actually involved
in the ‘static’ loading.
(c) In the last two decades, it has apparently been recog-
Figure 6.12 Utilising the velocity ratio (squared) or RQD%/100, nised in soil engineering that strain levels associated with
to estimate the ratio of the dynamic modulus for intact normal foundation designs are rather small, for example
samples (EL dyn), compared to the dynamic field modu- 0.01 to 0.1%, and therefore stiffnesses may be success-
lus (EF dyn). Deere et al., 1967. fully described by the correlations obtained from in situ
Deformation moduli and seismic velocities 107

seismic measurements (Matthews et al., 1997). Such


measurements also have the great advantage of register-
ing the stiffness of the ground at in situ stress levels and
in the undisturbed condition.
A corollary to the above is the predicted high deform-
ation moduli for jointed rocks at depth, shown in Figure
6.14 (Barton, 1995). A pre-condition here is that the
rock mass is undisturbed and strains are small, corres-
ponding more closely with the higher seismic velocities
seen at greater depths in the same rock masses. When a
tunnel or test adit is constructed at considerable depth,
the EDZ effect will alter the above conditions in a com-
plex way, to a degree that depends on rock quality and the
care with which the adit-excavation has been performed.
One must expect a certain seismic velocity gradient,
a deformation modulus gradient, a deformation gradi-
ent and even a permeability gradient and pore pressure
gradient, and finally a possible gradient of saturation.
(a) The natural complexity of a site may also tend to
increase the range of moduli and moduli ratios in rela-
tion to those measured in one particular lithology. The
Latiyan Dam site in Iran was founded on granites, peg-
matites, migmatites and gneiss, and weathered layers of
each of these. Lane, 1964, compares four of the basic
moduli commonly obtained at dam sites:

● laboratory EL dyn (from laboratory Vp, Vs and


Poisson’s ratio)
● field EF dyn (from field Vp, Vs and Poison’s ratio)
● field deformation modulus (D, Ed or M) (from
plate or flatjack loading)
● field elastic modulus (Ee) (from plate or flatjack
unloading)
(b)
and also gives ratios of each, as obtained in three
exploratory tunnels in the dam foundations. The surpris-
ingly large ratios of these moduli, for rather poor rock
mass qualities (Q mostly  0.1?) are shown in Table 6.3.
Link, 1964, also gave a wide-reaching comparison of
dynamic and static moduli from projects (usually dams)
in many countries. He found that the ratio of EF dyn/D
from field tests (category 6 in Table 6.3), ranged from
about 1 to 16, with most values of the ratio lying in the
range 3 to 7. A large number of the static loading tests
were pressure chamber tests from Central European
dam sites, and the author pointed out that the seismic
(c)
measurements (from Vp, Vs and dynamic Poisson’s ratio)
Figure 6.13 An extensive collection of ‘reliable’ Edyn. data, and of were also of quite large scale. As we see from the above
the ratios Ee/Edyn. and Ed/Edyn.. Stacey, 1977. (See table, ratios of EF dyn/D can be even larger than the above
individual references in article). when the rock quality is poor, due to the basic inequality
108 Rock quality, seismic velocity, attenuation and anisotropy

Figure 6.14 High values of static deformation modulus Emass (or M) for the rock mass (also referred to as D, or Ed by some authors), are pre-
dicted where stresses and rock quality are high. Here Emass approaches EF dyn.. Barton, 1995.

Table 6.3 Ranges of moduli and ratios of moduli at three complex sites (after Lane, 1964).

Modulus Tunnel 1 Tunnel 3 Tunnel 4

1 Laboratory Dynamic (EL dyn) (GPa) 35.8 45.4 35.8


2 Field seismic (EF dyn) (GPa) 17.7 25.4 16.4
3 Average modulus of deformation field loading (GPa) 1.8 1.3 1.3
(D or Ed or M, depending on author)
4 Average modulus of elasticity field unloading (Ee) (GPa) 4.8 4.6 3.8
5 Ratio laboratory dynamic to field seismic (EL dyn/EF dyn) 2.02 1.78 2.18
6 Ratio field seismic to modulus of deformation (EF dyn/D) 10 20 12
7 Ratio field seismic to unloading modulus of elasticity (EF dyn/Ee) 3.7 5.6 4.4
8 Ratio laboratory dynamic to modulus of deformation (EL dyn/D) 20 36 27
9 Ratio laboratory dynamic to jacking modulus of elasticity (EL dyn/Ee) 7.4 9.9 11.5

of the two static unloading/loading moduli: Ee/Ed Yugoslavian dam site tests. A general trend was noted as
(or Ee/D), shown, for example by Kujundzíc and follows:
Grujíc, 1966.
Link, 1964, made special reference to an extremely Edyn
Ee  (6.11)
high value of EF dyn measured (interpreted) under the (5.3  0.05Edyn )
lower slopes of the Vajon limestones. The value of
140 GPa was considered the result of high overburden, where Ee and Edyn are expressed in GPa.
and/or residual stresses. The lower the dynamic modulus the larger the ratio
Graphic presentation of the inequalities between of the dynamic/static moduli. When the inequality of
EF dyn and Ee were given by Kujundzíc and Grujíc, 1966. Ee and Ed is also considered, the very large ratios of
Figure 6.15a and b show the significant inequality of Edyn/Ed of 10 to 20 given by Lane, 1964, and Link,
Edyn and Ee for the case of the limestones tested in their 1964, are more readily understood.
Deformation moduli and seismic velocities 109

(a)

(b)

Figure 6.15 Inequality between Ee (static unloading) and Edyn. (or Figure 6.16 Average curves for the field-scale dynamic elastic (E:
EF dyn), for limestones tested at Yugoslavian dam sites. curve 1), bulk (k: curve 2), and shear (
: curve 3)
Kujundzíc and Grujíc, 1966. moduli, in relation to refraction seismic Vp, based on
80 examples from igneous and metamorphic areas.
6.3 Some examples of the three Sjøgren, 1984.
dynamic moduli
The authors found that the dispersion of the dynamic
Sjøgren et al., 1979, gave average curves for the three Poisson’s ratio values always increased at lower velocities
dynamic moduli (Young’s, bulk and shear), reproduced in (e.g., dispersion  0.02 at 5.5 km/s, and  0.065 at
Figure 6.16. These were based on 80 examples from three 3.5 km/s). In Figure 6.17, their calculated values of
igneous and metamorphic rock areas (1  dynamic E dynamic Poisson’s ratios (from equation 6.7) are shown
modulus, 2  bulk modulus K, and 3  shear modulus integrated with Vp magnitudes, with calculation of
G). Each are shown in relation to measured values of Vp equivalent values of Edyn and k bulk. The reduced mod-
from refraction seismic, which ranged from 3 to 5.5 km/s. uli where even higher values of dynamic Poisson’s ratio
Deviations from the curves occur with changes in the are calculated, can be readily imagined.
dynamic Poisson’s ratio (d). This is because of the fun- In this connection it is of interest to refer to the results
damental linkage between Vp and Vs: of Vp and Vs logging of shallow boreholes. Results given
by Chang and Lee, 2001 from a Korean tunnelling pro-
1 ject, show the elastic ‘constants’ for a 30 m deep borehole.
Vp  1    2

  d 
 (6.12) (Figure 6.18a). The Edyn moduli of 0.07 and 0.23 GPa
Vs  1 2  d  are less than the bulk moduli (k) of 0.15 and 0.38 GPa
in the residual soils in the top 10 m of the hole, where
and because of the elastic equations (6.5 to 6.10) linking the dynamic Poisson’s ratios are as high as 0.43 and
Vp and Vs with the dynamic elastic (Edyn), bulk (k bulk) 0.40 respectively, and Vp only 0.5 km/s.
and shear (
) moduli. Concerning Edyn and
, high It is not until the soft rock is reached at 24 m depth,
Poisson’s ratios give values below the curves in Figure that Edyn of 5.92 GPa becomes greater than k of 5.20 GPa,
6.16 and low Poisson’s ratios give values above the curves. with Vp increased to 1.85 km/s, and dynamic Poisson’s
110 Rock quality, seismic velocity, attenuation and anisotropy

Here we will make a brief, premature digression


to large-scale, high stress earthquake environments, to
illustrate one of the many uses of shear wave monitor-
ing. Chen et al., 1993, showed in research related to
earthquake source mechanisms, that stable sliding and
‘stick-slip’ could be differentiated by interpretation of
continuous shear-wave monitoring. Continuous sliding
between rock interfaces under pressure reportedly caused
a continuous decrease in shear wave amplitude, while stick-
slip behaviour was detectable by a rapid drop in ampli-
tude prior to or during ‘slip’, and by an increase in
amplitude during each ‘stick’ period.
The authors suggested this method as an alternative
to the classic but possibly unreliable observation that
the ratio of Vp/Vs showed a large decrease in earthquake
source regions prior to earthquakes (Agarwal et al., 1973).
(Anomalies in the amplitude ratio As/Ap have also been
reported before earthquakes.) A well known but tragic
fact is that Chinese seismologists succeeding in predict-
ing the 1975 Liaoning earthquake due to the decreased
Vp/Vs ratio, but failed one year later to predict the 1976
Tangsham earthquake which reportedly killed almost
one million people.
Figure 6.17 Dynamic moduli Edyn and k bulk, as a function of the In rock engineering projects, particularly arch dam sites
calculated dynamic Poisson’s ratio, for different Vp val- investigated by French consultants Coyne et Bellier,
ues. The solid curves refer to the appropriate magni- direct use has been made of the correlation between the
tudes of Edyn, and the dashed-curves refer to the k bulk ‘frequency of the second arrival’ (the S-wave frequency)
values. Note the reducing values of Edyn at higher values and the in situ static modulus of deformation (variously
of the dynamic Poisson’s ratios. Sjøgren et al., 1979. described as Ed D, Emass and M in the foregoing text
and referenced figures).
ratio reaching a ‘rock-like’ 0.31. With the harder rock at Schneider, 1967, and Londe, 1979, were instrumen-
30 m depth, the respective values have changed to tal in developing and publicising the ‘petite-sismique’
24.07 GPa, 18.43 GPa, 3.45 km/s and 0.28. method, which was subsequently given some impetus
Figure 6.18b shows a similar set of results for a deeper, by Bieniawski, 1978, with the additional test data and
mountain tunnel borehole, where Vp varied from 5.2 km/s seismic data from some South African projects. At this
beyond 30 m depth, to 5.5 km/s beyond 150 m depth. time a linear relation between in situ static modulus (D or
Here Edyn increased from 50 GPa to beyond 60 GPa, M) and S-wave frequency (f ) was proposed (D  0.054f-
with corresponding increases in rock density, and 9.2), as shown both in Figures 6.19 and 6.20.
reductions in Poisson’s ratio. Subsequently the method, termed SCARABEE
(Systeme Complet d’Analyse des Roches d’Appui de
Barrages et d’Excavations) was standardised and
6.4 Use of shear wave expanded by Coyne et Bellier and Geodia, 1995, in
amplitude, frequency and France. Use was made of a normalised value of the
petite-sismique Schmidt rebound (40), a normalised value of the
dynamic modulus (44 GPa) and of the S-wave fre-
In Chapter 5, we saw how S-wave frequency was used by quency (750 Hz), with a standard gain (45 dB), to
Stacey, 1977, to monitor the closure-under-stress of dry, give a more comprehensive picture of rock mass condi-
wet and clay-coated joints. P-wave amplitude was also tions. The correlation between D and f is now considered
found by Stacey 1977, Tanimoto and Ikeda, 1983, and to be of the non-linear form: (D  0.17 f 1.7 ), as shown
Tanimoto and Kishida, 1994, to be particularly sensi- in Figure 6.21, from Coyne et Bellier, 1998 and Carrère,
tive to joint closure and joint shearing mechanisms. 1999.
Deformation moduli and seismic velocities 111

(a)

(b)

Figure 6.18 The dynamic elastic moduli compared from down-hole Vp, Vs and density logging, a) for a shallow 30 m deep hole, b) for a
300 m deep mountain tunnel borehole. Chang and Lee, 2001. Note GPa, and symbol conversion, G to
.

6.5 Correlation of deformation and Q-deformation modulus relationships relate to large


moduli with RMR and Q scale data. The mean deformation modulus (Ed) data pre-
sented in Figure 6.4 for twenty different sites (as Ed versus
The extensive in situ deformation modulus data presented Vp2) is presented again in Figure 6.22, this time with mod-
by Ribacchi, 1988, which was reproduced earlier in this ulus in relation to RMR, shown as the lower scale.
chapter, represents a useful source of field data with which The source of the solid curve of Ed-versus-RMR given
to test how well the standard RMR-deformation modulus in Figure 6.22 is shown in Figure 6.23, where both the
112 Rock quality, seismic velocity, attenuation and anisotropy

Figure 6.20 ‘Petite sismique’ method of modulus estimation, with


(a) additional, linearly distributed data from South African
hydropower projects, from Bieniawski, 1978. Earlier
data from Schneider, 1967 is also shown.

(b)

Figure 6.19 The originally assumed, linear correlation between


S-wave frequency and tangent modulus of deform-
ation, in the so-called ‘Petite sismique’ method of Figure 6.21 Updated, non-linear relationship between the static
modulus estimation. Londe, 1979. modulus of deformation and the shear wave frequency,
from Coyne et Bellier, 1998, Carrère, 1999.

(Seraphim and Pereira, 1983) equation and the Q-system The Q-system based equation for deformation modu-
equation for ‘static’ deformation modulus prediction, are lus prediction gives a similar trend:
each shown.
1
It can be seen from Figure 6.23, that the Seraphim E d  M  10Q c 3 (6.14)
and Pereira, 1983, relation is as follows:
where Q c  Q  c/100, so Q c  Q in a typical
Ed  M  10((RMR 10)/40) (6.13) hard rock situation. This equation is the source of the
Deformation moduli and seismic velocities 113

‘circled-black-dots’ in Figure 6.22, which lie on, or just


below, the ‘RMR-curve’.
The solid and dashed lines shown in Figure 6.23
are also seen to almost coalesce for Q-values less than 1
and RMR less than 50. If we further assume that Q and
RMR can be related by the following simple relation
(Barton, 1995):

RMR  50  15 log10Q (6.15)

then we can supply a Q-value scale to the top of Figure


6.22, where in approximate terms:

RMR  5 20 35 50 65 80 95
Q  0.001 0.01 0.1 1.0 10 100 1000

For lower strength rocks, the estimate of static deform-


ation modulus is suitably lower using the Q-based
equation 6.14 compared to the RMR-based equation
6.13, and actually more in line with the extremely low
moduli that are usually measured, in the case of soft
Figure 6.22 RMR – Ed fit to Ribacchi, 1978 data using the empir- porous rocks. From earlier in this chapter one may note
ical Seraphim and Pereira, 1983 equation. The Barton, from Figure 6.7b (Kikuchi et al., 1982, Japanese tests)
1995 RMR to Q conversion: (RMR  15 log Q  50)
that one third of the deformation moduli lie in the
has been added to give the uppermost Q-scale, together
range 0.05 to 1.0 GPa, with correspondingly low P-wave
with ‘circled-dots’ representing the M  10Q c⁄ (GPa)
1
3

relation.
velocities (mostly 0.5 to 2.5 km/s).

Figure 6.23 A comparison of the old and the new RMR – deformation modulus relations (Seraphim and Pereira, 1983, and Bieniawski,
1989), and the more recent formulation involving Q or Qc. Barton, 1995.
114 Rock quality, seismic velocity, attenuation and anisotropy

An example application of equation 6.14 may be useful


here. Let us suppose that the conventionally logged Q-
value (for tunnel support selection) in a weak sandstone
of c  2 MPa is as low as 0.01 due to heavy jointing and
low frictional strength. When we correct Q  0.01 to
Qc, a value of Qc  Q  c/100, or Qc  0.0002 is
obtained. Substitution in equation 6.14 gives a predicted
Ed or M value of 0.6 GPa. The RMR equation 6.13 gives
a predicted Ed or M value of 1.8 GPa, when Q  0.01 or
RMR  20. Such an estimate may be too high for soft
rocks. As may be noted from Figure 5.36, in Chapter 5,
an additional correction for porosity is also available in
the ‘Q-diagram’, to improve fit to Tertiary type, weak,
porous sedimentary rocks.
In the case of an exceptionally poor quality faulted
zone or extremely weathered zone with Q  0.001, with
crushed rock pieces having a strength of only 1 MPa
(almost a soil) we would need to use a Qc-value of
0.00001 in equation 6.14, which would correspond to
an exceptionally low Ed or M value of only 0.2 GPa. This
(a)
is close to the ‘lowest possible’ M(min.) value tabulated in
Figure 6.14, and corresponds apparently (and perhaps
realistically) to a P-wave velocity as low as 0.5 km/s.
In Figure 6.24, showing Ribacchi, 1988 data from
Figures 6.5b and 6.6b, the following relation between

M and Vp is plotted. (This relation was given in Figure
6.14, see inset at top):

M  10.10((Vp 3.5) / 3)  10((Vp 0.5) / 3) (6.16)

The objective was to again test the relation against the


extensive data and data trends for limestones and
dolomites given by Ribacchi, 1988. In this figure, the
predicted M(mean) values given by this equation are seen
to represent upper-bound envelopes in relation to the
Ed data given in the figure. The M(min.) tabulation given
in Figure 6.14 appears to represent a lower-bound envel-
ope in the case of the five trend lines given in Figure 6.22.
It is important to try to understand why such high
velocities as, for example, 4.0 to 4.5 km/s can be associ-
ated with Ed values as low as 5 to 10 GPa as can be seen
in Figure 6.24b. Although the assumed porosities of the
limestones and dolomites might affect Ed values more
than Vp values, there is much evidence, for example
(b)
from Figure 2.10 (Chapter 2, Fourmaintraux, 1975)
that this cannot be so. Figure 6.24 Plate load data for limestones and dolomites (see Figs.
Possibly the most likely explanation, which has been 6.5b and 6.6b) from Ribacchi, 1988, and a compari-
raised earlier in this chapter, is that the large scale deform- son of the M-Qc model. (Mmean is the solid line: left,
ation modulus test as practised at dam sites (plate load, Mmin. is the dashed line: right).
Deformation moduli and seismic velocities 115

flatjack or occasionally pressure chamber) is nearly from core logging and refraction seismic, and from
always testing an excavation disturbed zone (EDZ), in deeper cross-hole tomography data. The second equa-
a loading direction parallel to the radial stress ( r), while tion was developed from large scale modulus of deform-
the velocity measurement may be averaging velocities ation data where either the Q-value was known from
over a larger volume, and may tend to be recording independent logging or from approximate conversion
velocities parallel to the (tangential stress) direction, from logged RMR values, using equation 6.15. The
which is a much higher, maximum local principal stress, implied connection between M and Vp (elimination of
compared to the minimum radial stress. The latter obvi- Q c between the above equations) to give equation 6.16,
ously approaches zero, due to the effect of excavating a is a ‘pure’ link, which ignores the potential complica-
test adit. tion that anisotropic stresses (  r) may cause
The stress anisotropy (  r) around an excav- anisotropic velocities in the EDZ test zones.
ation, which can explain higher values of Vp parallel to
than parallel to r, could therefore explain the higher-
than-expected Vp values, and the higher-velocity-than-
expected Ed–Vp trend lines seen in Figure 6.24b.
In this connection, it is significant that the empirical
equations that were introduced in Chapter 5:

1
Vp  3.5  log10 Q c and M  10Q c 3

were developed independently (at different times) from


different data sets. The first equation was developed
7 Excavation disturbed zones and
their seismic properties

The existing surface of exposed rock that may be found 7.1 Some effects of the free-surface
where soil cover is absent can be considered as nature’s on velocities and attenuation
disturbed zone. Steep mountainsides in rocky terrain,
steep gorges or valley sides where dams may be founded The problem with the EDZ is that many of the geome-
or glaciated terrain are typical examples. As we have seen chanical properties we are most interested in investigat-
earlier the special coincidence of low stress, weathering ing at large scale are themselves affected by the process
effects and possibly more frequent jointing (with joint of obtaining access. Only a borehole-size intrusion may
apertures above the limits of ‘acoustic closure’) make such be considered nearly free of damage, but it will cause local
zones extremely ‘visible’ in seismic refraction or sonic disturbance to the stresses, that may influence permeabil-
logging of shallow boreholes. Velocities may be up to sev- ity measurements and may also influence preliminary
eral km/s lower in the near-surface zone than at greater load cycles from borehole dilatometers. The ‘skin-effect’
depth, in some cases even when joint frequencies remain around an excavation adit is far from ‘skin-deep’, and
unchanged. In an attempt to get away from the wea- may extend many metres or to one or several radii from
thered zone, investigators of dam sites and deep founda- the excavation wall, especially in the case of softer rock
tions may often construct adits or shafts for conducting that is highly stressed.
deformability and other geomechanical tests. Unfortu- We must assume both a damaged zone and a disturbed
nately, however much care is taken, even to the extent zone that together have been given the nomenclature
of non-blasting methods, a disturbed zone results. The EDZ (excavation damage and disturbed stress zone).
removal of stressed rock and its usual replacement by air ED1D2Z would be more specific, as with e.g. line drilling
at atmospheric pressure (a convenient ‘definition’ of tun- of an experimental tunnel, as at URL in Canada, we have
nelling), results in a radial stress ( r) that approaches only ED2Z.
zero at the excavation walls. The tangential stress ( ) An unusual and instructive geophysical monitoring of
may assume many values (including negative) depending the Dneiper ship lock excavation, which reached a depth
upon the existing stress anisotropy, joint orientation, rock of more than 20 metres, is shown in Figure 7.1. The effect
strength and the disturbance caused by the excavation of loosening caused by blasting, stress relief (and pre-
method. sumably inadequate slope reinforcement) is shown very

Figure 7.1 Free-surface effect, and slope excavation (and degradation) effects on P-wave velocities. There is a 1 year delay between measure-
ments c) and d). Savich et al., 1983.
118 Rock quality, seismic velocity, attenuation and anisotropy

clearly. There is a one-year delay between diagrams (c) seismic probe for easier access to inaccessible locations
and (d). Savich et al., 1983, refer to a 200–300% reduc- behind steep slopes. Cross-hole measurements clearly
tion in velocity, a 75 to 85% reduction in deformation indicated the presence of stress-relieved joints behind
modulus, and a 1 to 20 times increase in ‘joint voids’. the slope faces.
This can be considered an excavation disturbed zone The excavation efficiency of large dragline excavators in
(EDZ) study related to slopes, in which velocity open cast coal mines as a function of efficient blasting
changes were related both to loosening and presumably practice was investigated by Young et al., 1985, using
to water drawdown. The same basic mechanisms may be what they called ‘seismic spectroscopy’ to quantify the
at work around tunnels, in which EDZ effects are import- degree of brokenness of the rock before and after blasting.
ant for may reasons, including increased support needs, They utilised seismic attenuation and anisotropy dia-
reduced deformation modulus, and frequently increased grams to evaluate the efficiency of the blasting process.
leakage (or permeability) in the case of larger or less sta- Certain frequencies were found to be attenuated due to
ble tunnels. These effects may be remedied by high pres- the effect of fracture size and porosity (and void space).
sure pre-grouting with micro or ultra-fine cements. This The principle of the method is shown in Figure 7.2.
will be discussed later. In the case of very stable rock, and The attenuation spectrum of the broken rock mass
small tunnels, inflow may reduce in relation to expect- differed fundamentally from that of the unbroken, more
ation. (Olsson, 1992). homogeneous rock. As shown in this most interesting
Price et al., 1970, reported the results of seismic refrac- Figure 7.3, attenuation was larger and more irregular for
tion measurements at three sites in the UK where bolt-
ing, anchoring and grouting of rock slopes was being
designed. Their measurements at Edinburgh Castle in
complexly jointed basalt foundations, suggested a 16%
increase in Vp as a result of bolting and anchoring (6 to
12 m deep) followed by grouting.
At Cheddar Gorge, which cuts through massive, but
bedded limestones, an outer layer of about 5 m thickness
showed a velocity of 2.9 km/s, while deeper into the walls
of the gorge, the velocity was 5.5 km/s.
The vertical, pseudo columnar jointing in dolerite at
Stirling Castle showed ratios of Vfield/Vlab as low as 0.1
due to stress relieved joints and weathering near the sur-
face. Slope ‘EDZ’ can clearly exceed the EDZ associated
with tunnelling, due to the added influence of weathering.
The authors also described early trials with a downhole

Figure 7.3 Effect of blasting and depth below surface on the seis-
Figure 7.2 Seismic attenuation compared in broken rock, and in mic attenuation. Attenuation is greatest near-surface,
compact rock masses. Young et al., 1985. and at intermediate frequencies. Young et al., 1985.
Excavation disturbed zones and their seismic properties 119

shallow depths (where fractures open more easily) and authors cited the same reasons for the velocity reduc-
more regular and limited at greater depths, where con- tion that we hear in more recent times at nuclear waste
finement limits fracture or joint opening. Associated investigation sites such as Hanford (King et al., 1984),
compressional wave velocities typically changed from URL (Maxwell and Young, 1996), Äspö (Emsley et al.,
1.1 and 1.8 km/s (at the surface/at greater depth) prior 1996) and Stripa (Olsson, 1992). Fracture formation,
to blasting, down to 0.6 and 1.1 km/s respectively, after joint disturbance, stress redistribution and possible des-
blasting. iccation of the existing joint system were all listed by
These excavation disturbance effects were accom- Hasselstrøm et al., in 1964, and are equally relevant (and
panied by rotations of both attenuation anisotropy axes complicated) today.
and velocity anisotropy axes, presumably as a result A classic EDZ investigation in relation to pressure tun-
both of disturbance to pre-existing joint patterns and nel design was reported by Kujundzíc et al., 1970. They
blast-induced fracturing in the proximity of a high wall. performed a trial chamber test for investigating post-
stressing effects on the concrete liner of their 5 m diam-
eter, circular tunnel. In the course of this study, they
7.2 EDZ phenomena around utilised numerous grouting boreholes (32 in all) for
tunnels based on seismic conducting cross-hole seismic along the tunnel axis at
monitoring eight different radial positions. Their results are shown
in Figure 7.5.
In the years before the 1980s, reports of EDZ investi- They visualised the existence of three zones around
gations in the rock mechanics literature were usually in the tunnel: 1) the loosened zone (with lowest velocities);
connection with hydropower projects. An impressively 2) the stress bearing ring (with highest tangential stresses
early model for subsequent investigations was provided and velocities); 3) the uninfluenced zone (with declining
by Hasselstrøm et al., 1964, at a dam site in Sweden, who velocities and background stresses). Their mean results
compared cross-hole and downhole sonic logging results (Vp  3.5 km/s at the tunnel wall, Vp  5.5 km/s at 1 m
in an investigation gallery, as shown in Figure 7.4. radius, and Vp  4.5 km/s in the undisturbed zone)
Velocities were seen to fall from about 5.5 to 3.5 km/s shown in the centre of Figure 7.5 can be interpreted by
in the outer 1 metre of their 1.5  2 m gallery. The means of the Vp-stress effect model as discussed in
Chapter 5 (Figure 5.36).
Significant sophistication was added to the analysis
of disturbed zone phenomena by Russian engineers,
who analysed a variety of effects, including anisotropic
velocities using ultrasonic and seismic methods. They
emphasised the need to consider the use of different
wave lengths (see Chapter 3, Lykoshin et al., 1971,
Figure 3.12).
In relation to the EDZ logging of a shaft in diabase,
these same authors used a time-average equation to esti-
mate the joint void ratio (e) as a function of depth, as
shown in Figure 7.6:

Va ( Vr  Vm ) (7.1)
e
Vm ( Vr  Va )

where Va was given the value 345 m/s assuming air-


filled joint voids, Vr is the intact rock velocity and Vm
Figure 7.4 An early EDZ measurement at a Swedish dam site gallery the mean velocity at the depth of measurement. For
that compares cross-hole (1) and sonic log (2) results. example, we can substitute the P-wave velocities 0.345,
Hasselstrøm et al., 1964. 5.0 and 3.5 km/s respectively to obtain an estimated
120 Rock quality, seismic velocity, attenuation and anisotropy

Figure 7.5 Pressure tunnel investigations of seismic velocity as a function of radius, and therefore as a function of tangential stress, using
cross-hole measurements. Kujundzíc et al., 1970.

joint void ratio of 0.03 (or 3%) at an imagined shallow stress) cause Vm eventually to equal Vr, then clearly the
depth into the wall of an excavation. void space will theoretically vanish. Lykoshin et al.,
At greater depth when Vm has increased, for example, 1971, recommended the use of ultrasonic logging with
to 4.5 km/s, the calculated joint void ratio will have the minimum transmitter-receiver distance (e.g., 0.1 m)
reduced to 0.008 (or 0.8%). If we suppose that greater to obtain the relevant velocity distributions in such
depth into the excavation wall (or the results of high studies.
Excavation disturbed zones and their seismic properties 121

Figure 7.6 Calculations of joint void ratio (%) in the EDZ sur-
rounding a shaft in diabase, using a time average equa-
tion. Lykoshin et al., 1971.

The simple time-average equation given above is a par-


ticularly useful theoretical means of investigating the void Figure 7.7 Tunnel EDZ measurements assembled by Capozza,
space created in an EDZ, since we can also investigate the 1977 for a variety of rock types. Tunnel depths (m) are
shown in parentheses.
theoretical effects of drying out, by substituting the Va
value of 0.345 km/s (as in equation 7.1) in place of a Vw
value for water of 1.44 km/s. However, the resultant val- Values in parentheses show the depth of the tunnel in
ues of Vr for the intact rock and Vm for the jointed rock metres. It is typical for Vp to be reduced by 25 to 50%
will also tend to change due to drying, most rapidly in the close to the tunnel walls. The 3 m thick ‘skin’ observed
case of Vm (because of the influence of Va) and in the tunnel with 2100 m of overburden is presumably
less rapidly in the case of Vr. a result of stress slabbing in the granite.
The example of reducing Vp with time due to grad- Bernabini and Borelli, 1974, described a variety of
ual drying of an intact sample of granite (Nur and early seismic tests performed at hydroelectric projects in
Simmons, 1969, Figure 2.18, Chapter 2) could be ref- Italy in the 1950s, 1960s and early 1970s. In hydropower
erenced at this juncture. Lykoshin et al., 1971, also gave tunnels in gneiss and granite, they measured EDZ effects
a more complex expression for the velocity of eventual using seismic refraction techniques along four lines in
joint filling materials, based on the P-wave velocities Va each tunnel (two in the arch, and two in the lower walls).
and Vw, and on three other component velocities of the There were marked reductions of velocity (at least 50 to
rock and rock mass. 60%) in the measurements made just 100 metres from
Capozza, 1977, reviewed the results of a large num- the tunnel entrance (for example 4.8 to 2.2 km/s) as can
ber of EDZ-style investigations of tunnels. These were be seen in Figure 7.8. Further from the tunnel entrance
made using both seismic and cross-hole seismic tech- (300 m and 500 m) the higher quality and higher vel-
niques. Figure 7.7 shows the ratio of velocities V2/V1 in ocity rock showed less marked reductions in velocity due
the first few metres depth in numerous tunnel walls. to excavation disturbance.
122 Rock quality, seismic velocity, attenuation and anisotropy

Dhawan et al., 1983, described seismic refraction the location of minimum tangential stress, if horizontal
measurements at a dam site in the Himalayas and slots, rather than vertical slots were cut in the wall, and
attempts to correlate results with 1100 metres of core log- if horizontal along-the-valley stresses were maximum
ging and in situ testing in the same quartzitic and slaty principal stresses in the location of the tests.
phyllites. The in situ testing was performed in boreholes The disturbed zone surrounding the drift showed a
and in an investigation drift in the dam foundations, thick low velocity zone (1.4 km/s) of about 3 m
shown in Figure 7.9. thickness in the first 10 metres of the drift. This
The values of deformation modulus, ranging from exceeded the diameter of the drift (2 metres). Further
about 0.5 to 3 GPa from plate jacking and flat jack tests, inside the drift, in less weathered material, the higher
seemed to correlate poorly with the wider range of seismic velocity material (Vp  2.85 km/s) came closer to the
velocities (1.5 to 4.0 km/s), possibly due to the discrep- walls of the drift, as shown in Figure 7.9.
ancy of the plate jacking loading direction (radial) in rela- In the same year, Bonapace, 1983 described rock
tion to the perpendicular-to-tangential-stress direction mechanics testing for the design of an extremely high
followed by most of the seismic refraction ray-paths. The head (1260 m) inclined pressure shaft. Deformation
flat-jack loading could possibly have been performed in modulus testing at different scales culminated in a radial

Figure 7.8 Seismic refraction profiles along tunnels in gneiss and granite give a crude indication of EDZ. Berabini and Borelli, 1974.
Excavation disturbed zones and their seismic properties 123

jacking test chamber (diameter  3 metres) which gave Downhole logging by Stead et al., 1990, in a potash
the lowest modulus of 11.5 GPa. In order to extrapolate mine in Saskatoon, Canada also showed distinct EDZ
this result along the shaft alignment, 4 m deep boreholes effects as shown in Figure 7.11a and b. Holes in the
were drilled at 10 m intervals, in order to measure seismic roof showed velocity anomalies connected with the
velocities and correlate these with the measurements at opening of clay seams (or the presence of clay seams),
the radial jacking location. while anomalies and reductions in velocity in the walls
These EDZ-style measurements in the 4.8 m diameter of mine entries were reportedly associated with stress-
shaft wall, shown in Figure 7.10, gave velocities as low as 1 induced fracturing of the potash. Freshly mined open-
to 1.5 km/s in the outer 1.0 metre of the drill-and-blasted ings showed only slight reduction in velocity in the
shaft, but velocities no lower than 2 km/s in the outer 0.5 future EDZ zone, as fracturing took time to develop.
metre of the TBM excavated section. In general, velocities The P-wave velocity clearly appeared to be more sensi-
were 1 km/s higher in the case of the TBM excavation, tive to these anomalies than the S-wave velocity, judg-
over the depth range of 1 to 4 metres from the shaft wall. ing by the two figures.
Holcomb, 1988, found that mining of excavations in
the bedded salt of the WIPP site in New Mexico had
greatest effect on the attenuation of compressional waves.
The EDZ could be detected to a depth of some 3 metres
(about 1/2 D). The mining induced radial stress relief and
increased tangential stress would tend to cause under-
saturation due to dilation of the salt. The reduction in
compression wave amplitude, which increased with time
after excavation, was a stronger indicator of EDZ than the
minor reduction in P-wave velocity (0.5 km/s).
In softer rocks such as chalk the EDZ effect can be even
more marked in terms of percentage change in velocity.
At a storage cavern site in Eocene chalk with one pre-
dominant set of vertical joints of about 1 m spacing and
a second, less well developed set, McDowell et al., 1992,
showed mean background velocities of 2.34 km/s redu-
cing to 1.47 to 1.56 km/s in the outer 1 to 2 metres of the
Figure 7.9 EDZ effect accentuated by the initial weathered zone wall of the pilot tunnel.
at an investigation adit at a dam site in the Himalayas. Borehole investigations at the same site had shown
Dhawan et al., 1983. significant increases with depth; Vp was 1.25 km/s from

Figure 7.10 Contrasting EDZ effects on Vp from drill-and-blast and TBM excavated shaft. Note the travel time-distance-velocity plotting
format. Bonapace, 1983.
124 Rock quality, seismic velocity, attenuation and anisotropy

RQD variations. This suggests a stress-related increase in


Vp, rather than systematically reduced jointing. In this
relatively low velocity chalk, the authors experienced the
seismic restriction caused by shotcreting of a fault zone.
This stiffer, high strength layer caused a thin, high vel-
ocity surface layer to be registered.

7.3 EDZ investigations in


selected nuclear waste
isolation studies

Starting in the late 1970s, but increasingly from the 1980s


and since that time, there have been a growing number of
nuclear waste disposal studies at potential or generic sites,
and at underground laboratory facilities. Some of the
facilities have become ‘household’ names in rock mechan-
ics circles, starting with Stripa (in Sweden), Grimsel (in
Switzerland), URL – underground research laboratory (in
Canada), BWIP – basalt waste isolation plant (in the
USA), Äspö (in Sweden) and several other prominent
facilities. In this review of selected EDZ data, significant
results from four of these facilities will be described here.
In Part II, Chapter 16, further results from additional sites
in the United States will be described, due to their seismi-
cally-relevant jointed-rock-mechanics aspects.

7.3.1 BWIP – EDZ studies

King et al., 1984, and King et al., 1986, reported on a


series of cross-hole seismic measurements performed in
a flow-entablature, columnar-jointed basaltic rock mass.
The columns were regular but sinuous, 0.15 to 0.36 m
in thickness, dipping 70 to 90°, with frequent low angle,
discontinuous cross-jointing. The measurements were
Figure 7.11 EDZ effect and clay seam velocity anomalies in a made between four horizontal boreholes drilled 12 metres
potash mine roof. Vp is seen to be more sensitive to into the wall of a drill-and-blasted underground opening,
the EDZ than Vs. Stead et al., 1990. at 46 m depth, and theoretically above the water table.
The objective was to investigate the effect of blast dam-
age and stress redistribution, i.e., two of the assumed chief
0–30 m, about 2.2 km/s at 70 m depth, and 2.6 km/s at components of the EDZ or excavation damage and
90 m depth. disturbed zone.
(This closely follows the Q c  0.1 velocity–depth curve A vertical separation of boreholes 1 and 2 of 3 m, and
in Figure 5.37, with a 10% porosity correction, but a horizontal separation of boreholes 3 and 4, also of 3 m
requires an upward correction to an appropriate rock (in the form of a cross) allowed both the vertical, hori-
quality Q-value: e.g. to Q  1 if c was 10 MPa.) zontal and diagonal paths to be investigated, thereby
A corresponding increase in S-wave velocities was not crossing different joint sets (predominantly sub-vertical
recorded in the chalk; values were mostly 1.4 km/s, with columnar) at different angles. Figure 7.12a and b show
small variations (0.3 km/s) related to lithology and the test set-up.
Excavation disturbed zones and their seismic properties 125

(a) (b)

(c) (d)

Figure 7.12 Experiment for measuring EDZ effects in a tunnel wall (or ‘face’), at the Basalt Waste Isolation Project (BWIP). a) Test set-up,
b) test principle, c) and d) four of the selected ray paths and the effect of the EDZ on the P-wave velocities. King et al., 1984.
126 Rock quality, seismic velocity, attenuation and anisotropy

Figure 7.14 Effects of stress on laboratory samples of the basalt, in


Figure 7.13 RQD variation in the boreholes in basalt showed no the dry and saturated states. King et al., 1984.
marked EDZ effect. King et al., 1984.

than in the intact laboratory samples (6.3 km/s compared


The large contrasts in Vp values for the vertical path to 6.0 km/s). In the dried-out state, the parameters show-
(#1 to #2) and for the horizontal path (#3 to #4) close to ing most response to stress level were Vp and the dynamic
the opening (1.5 to 2.0 km/s difference) are indicated in Young’s modulus.
Figure 7.13c. A significant reduction in velocity of 1 to The authors also analysed in some detail the seismic
2 km/s, most in the diagonal directions shown in Figure quality factor which also has the symbol Q. This seismic
7.13d, is seen in the outer 2 to 3 metres. There is also quality (which we have termed Q seis to avoid confusion)
some indication of a tangential stress concentration is inversely proportional to the attenuation coefficient,
effect; the background (far-field) velocity of about 5.4 for a given frequency. Only in the vertical direction was
to 5.8 km/s appears to be elevated by about 0.5 m/s it significantly increased in situ, perhaps in response to
from about 4 to 8 m depth in the wall. the tangential stress increase. King et al., 1986 laboratory
Signs of loosening effects of the sub-vertical joints are test data for the same rock also showed the positive effect
evident in the horizontal velocity reductions. The add- of stress on Q seis, with the S-waves showing the strongest
itional effects of potential stress anisotropy and joint reduction in attenuation and the highest increase in Q seis,
anisotropy are clearly illustrated in King et al., 1984, 1986, as clearly shown in Figure 7.15.
results. The authors registered no consistent trend in RQD In a subsequent paper, Zimmerman and King, 1985,
values with depth. Figure 7.13 does not suggest that used the known effects of saturation on dynamic moduli.
increased jointing caused the marked velocity reduction. Theoretically, shear modulus is unaffected, while bulk
The velocity reductions are a product of blast-damage, modulus increases linearly to equal the uncracked rock
stress relief (and redistribution) and possible reduction in when s  100%. They suggested that the degree of
moisture content. The authors noted water flow from saturation and the joint or crack density were each con-
some of the horizontal holes during the tests and originally tributing to the measured reductions in cross-hole
assumed more or less saturated conditions. The horizontal velocities Vp and Vs, as shown in Figure 7.16a and b.
and (to a lesser extent) the diagonal seismic measurement These results are presented in order to emphasise the
paths crossed the more open columnar joints, and these possibility of drying out of some of the joints, despite
features clearly opened most as a result of excavation, giv- water flow from some of the holes. The theoretical analy-
ing the strongest reductions in velocity (55% to 65%). sis of crack density does not appear to be supported by
King et al., 1986, also investigated the effects of stress the RQD measurements in general (Figure 7.13) but is
(7 to 21 MPa) and saturation on Vp, on Vs, on the perhaps an expression of joint void ratio changes (equa-
dynamic modulus, and also on the Poisson’s ratio of ten tion 7.1) with the joints closest to the tunnel wall show-
intact specimens. Comparison of Figure 7.12 and Figure ing the largest voids and therefore suggesting an apparent
7.14 indicate a higher velocity in the vertical in situ planes (but false) increase in joint density.
Excavation disturbed zones and their seismic properties 127

7.3.2 URL – EDZ studies

The underground research laboratory (URL) in Mani-


toba, Canada was the subject of numerous and very
informative rock mechanics and geophysical studies.
Novel instrumentation was used for monitoring responses
to test tunnel excavation through the limited areas of nat-
ural fracturing. However, the dominance of massive,
unjointed, highly stressed granite resulted in particular
focus on stress-related EDZ, with down-hole sonic log-
ging and acoustic emission monitoring, together with
parallel laboratory tests and numerical modelling studies.
Several phases of investigation of EDZ effects, and several
scales of investigation were accomplished in this mostly
massive, highly stressed granite. (Martin et al., 1995).
We can start at the smallest scale, by looking at the
effect of the high in situ stresses, and mode of excavation,
on the state of micro-cracking in cored samples. The
French group, ANDRA (Homand-Etienne and Sebaibi,
1996) selected core samples from radial boreholes drilled
in the walls of a drift where excavation was either by nor-
mal blasting or by smooth blasting. Micro-cracking, partly
induced by the excavation and partly pre-existing, indi-
cated Vp reductions of about 1 to 1.5 km/s in the outer
0.8 m of the normally blasted tunnel, and reductions of
about 0.5 km/s in the outer 0.5 m of the smooth blasted
excavation, as shown in Figure 7.17. In each case the core
samples recovered from various depths into the tunnel
wall had been machined into cubic specimens prior to
Figure 7.15 The effect of stress on the seismic quality Q, the
the velocity measurements.
inverse of attenuation, for the basalt samples. Note
The effects of highly anisotropic, sub-horizontal stresses
that the authors’ use of the term ‘rock quality factor’
is left unchanged in this reproduction of their draw-
at the URL were studied in a unique test tunnel (ED2Z
ing. (As will be discussed in detail in Part II, Chapter only), at the 420 m level. (Figure 7.18). Excavation was
10, there are indeed certain numerical similarities by line drilling and reaming, followed by mechanical
between rock quality Q and seismic quality Q). King breakout to avoid blast damage. Principal stresses of
et al., 1984. approximately 60, 45 and 11 to 15 MPa, caused classic
‘break-out’ resembling that in a borehole.
The originally intended circular, 3.5 m diameter tun-
nel, was excavated parallel to 2 in order to maximise
the potential for stress-induced fracturing. The isotropic-
elastic theoretical tangential stresses of 165 MPa
(3 1 – 3) in the ‘rotated’ 11o’clock roof and 5o’clock
floor, and 15 MPa (3 3 – 1) in the side walls, caused
prominent V-shaped notches of rock failure. Associated
micro-cracking and stress changes were imaged tomo-
graphically (Maxwell and Young, 1996) as will be shown
Figure 7.16 Vp and Vs as a function of depth in the tunnel wall (or shortly.
‘face’), saturation level and crack density variations in The walls of boreholes drilled in this carefully excav-
the EDZ in BWIP basalt. Zimmermann and King, ated ED2Z, provided important information about
1985. stress-related disturbance. This was measured directly
128 Rock quality, seismic velocity, attenuation and anisotropy

in 1 m deep boreholes using a micro-velocity probe


with 10 cm separation of the transducers. Maxwell and
Young, 1996, installed their probe in holes drilled both
into the tensile region in the test tunnel wall, and into
the compressive region. Since their measurements of
velocity were in a radial direction, micro-cracking was
registered in the ‘fast’ direction and in the ‘slow’ direc-
tion, the latter perpendicular to the micro-cracking.
Figure 7.19a shows how both the P-wave and S-wave
velocities reduce by some 10% in the outer 1 m of the
tunnel wall in the tensile region of tangential stress. The
unstressed core from the same hole showed additional
damage and lower velocities due to the loss of radial
stress (Figure 7.19b).
Carlson and Young, 1993, and Maxwell and Young,
1996, also recorded acoustic emission (AE) locations dur-
ing careful (mine-by) extension of the 420 m deep test
tunnel. Calculated compressional wave velocities showed
quite strong anisotropy in this massive granite. This was
obviously caused by the principal stress anisotropy (60, 45
and 11 to 15 MPa). The lower hemisphere Vp stereonet
shown in Figure 7.20 gives a clear indication of the link
between stress anisotropy (caused by the very different
maximum and minimum tangential stresses), and P-wave
velocity. The authors’ velocity data showed a clear EDZ
Figure 7.17 Core samples recovered from 3 m long radial holes at
effect in the first 1 metre depth of tunnel wall with veloci-
a drift in the highly stressed URL, where both smooth-
ties reducing from about 5.8 km/s to 5.3 km/s in the case
blasting, and normal blasting were used, show evidence
of micro-cracking having originated in the differently
of the vertical direction. The acoustic emission results
configured EDZ. Homand-Étienne and Sebaibi, 1966. confirmed that rock failure was initiating just inside the
tunnel wall, orthogonal to the 1 direction. The authors

(a) (b)

Figure 7.18 A unique line-drilled and hand-mined test tunnel at URL, showed classic break-out related to the sub-horizontal 1 orientation.
Maxwell and Young, 1996. Note location of AE sensors.
Excavation disturbed zones and their seismic properties 129

considered that the Vp anisotropy (10 to 14%) was due to


open-crack porosity in addition to the micro-crack fabric.
The location of the microseismic sensors in relation
to the drilled-and-hand-mined test tunnel can be seen in
Figure 7.18. The AE events shown in Figure 7.21 are seen
to cluster both where tangential stresses were highest
and where seismic velocity (Vp) gradients were steepest.
Relatively decreased velocities were seen in the two
regions that were under tensile tangential stress.
It is of particular interest to note the ‘broadness’ of
the high velocity regions, which presumably reflect an
increase in deformation modulus due to the particular
alignment of the maximum tangential stresses. The
relevant isotropic elastic stress distribution calculated with
(a)
a boundary element program is also shown in Figure 7.21.
In subsequent work at the URL for an experimental
tunnel sealing experiment, Young and Collins, 1999
were able to demonstrate AE-interpreted reductions
in average P-wave and S-wave velocities in the highly
stressed zones caused by post-excavation of larger diam-
eter ‘dog-collars’ (or sealing-bulkheads) for forming sep-
arate concrete and bentonite seals. The particular tunnel;
Room 425, was elliptical in shape to reduce the previously

(b)

Figure 7.19 EDZ effect registered by downhole sonic probe at


URL test tunnel. a) Vp and Vs in situ down a 1 m deep Figure 7.21 AE events, and the interpreted regions of increased
horizontal hole b) P-wave velocity in situ compared average velocity, are reasonably consistent with the elas-
with that of unstressed, but micro-cracked core from tic continuum model of the stress distribution. Maxwell
the same hole. Maxwell and Young, 1996. and Young, 1996.

Figure 7.20 Principal stress-orientated velocity anisotropy (lower hemisphere plot) and EDZ effects on Vp at the same URL test tunnel, Canada.
Carlson and Young, 1993.
130 Rock quality, seismic velocity, attenuation and anisotropy

referred maximum tangential stresses to a tolerable and Vs. When constant density was assumed, these
105 MPa. increases suggested a theoretical 3.5% per year increase
The microseismic instrumentation available to moni- in Edyn for the concrete.
tor this tunnel sealing experiment (TSX) was extensive, The focus on microseismic and acoustic emission
consisting of 16 grouted-in-borehole triaxial acceler- monitoring when excavating variously shaped tunnels
ometers, operating in the 0.1–10 kHz band, and pro- in these highly stressed, massive granites has naturally
viding 3D coverage of an outer 100 m  100 m  50 m aroused curiosity about both the exact location of fail-
volume surrounding the tunnel. Two higher resolution ure around an opening (how close to the wall is initi-
AE arrays for recording in a 10 m  10 m  10 m vol- ation?), and curiosity concerning the type of fracturing
ume around the collars, consisted of 24 ultrasonic responsible for the microseismic events.
transducers operating in the 50–250 kHz band. Each of Cai et al., 1998, together with mining-rock-mechanics
these arrays had 16 receivers and 8 transmitters. colleagues Kaiser, and Martin, who had most responsi-
Young and Collins, 2001 described the way their AE- bility for URL rock mechanics experiments, collectively
based monitoring of temporal changes in Vp and Vs was address the dominance of tensile fracturing in brittle
used to estimate the theoretical change in crack density rocks, and the apparently unrealistic source sizes provided
(c) and saturation (s) along any particular ray path. They by shear-failure based models.
also refer to Maxwell et al., 1998 who used these tech- Contrary to failure around tunnels or boreholes in
niques to monitor (c) and (s) along radial boreholes at weaker, less brittle materials, which seems to be dom-
the line-drilled mine-by test tunnel shown in Figure 7.18. inated by log-spiral-type single-set or conjugate-set shear
The theoretical assumption was that the saturation (s) failure (e.g. Addis et al., 1991, Barton, 2004), the rock
did not affect the dynamic shear modulus (
), but that mass fracture process near underground openings in
the dynamic bulk modulus (k) increased linearly with hard brittle rocks appears to be dominated by extension
(s), even equalling that of uncracked rock when s  1. cracking, as extensively reviewed by Cai et al., 1998.
Concerning the crack density (c), defined as the num- The authors therefore argue that the focal mechanism
ber of cracks per unit volume multiplied by the cube of will differ from that of natural earthquakes, where slip
the mean crack radius, they assumed that the dynamic on pre-existing faults may dominate.
Young’s modulus Edyn and the dynamic Poisson’s ratio in Out of the 3,500 events recorded when excavating
the damaged material, when normalized by the undam- the line-drilled test tunnel depicted in Figure 7.18, some
aged material, decreased exponentially with (c). Young and 800 events were located within the volume of a subse-
Collins, 2001 showed a small reduction in (s) from 0.485 quently excavated 3.5 m thick slice, that was line-drilled
to 0.455, and a small increase in crack density (c) from from the floor of the tunnel. These events mostly clus-
0.198 to 0.206, during a 16 months monitoring period, tered in the region, shown in Figure 7.22a, where 2D
for a ray-path from the tunnel perimeter to 3.5 m into deviatoric stress (actually a shear stress) was larger than
the wall of the tunnel. the crack initiation threshold, which the authors sug-
The development of a rock fracture adjacent to the gested was about 70 MPa, or (0.4  0.1) c. This was
heat-producing curing of the concrete part of the bulk- also where the ‘notch’ failures occurred.
head, and its effect on AE monitoring of velocities, is also When the authors estimated the crack sizes using trad-
described by Young and Collins, 2001 and Young and itional shear models, they appear to have obtained unre-
Baker, 2001. This fracture caused a temporary loss of AE alistically large crack radii. They argue that the ratio of
signal, both for average P-wave velocity and P-wave the S-wave energy to the P-wave energy is an important
amplitude monitoring. There was subsequent recovery indicator of the type of focal mechanism. The S-wave
of both signals caused by remedial grouting. The energy usually dominates, as the energy radiated in
approximate 90% drop in P-wave amplitude caused by P-waves is only a small fraction of that of S-waves. It
the fracture was about 30% recovered by the effect of appears from their review that when Es/EP  10, the
grouting. cracking process involves a dominant tensile failure com-
A steady, slow rise in Vp (40 to 50 m/s increase was ponent, whereas if Es/EP  20 to 30 the cracking process
noted during one of the monitoring periods), was associ- is dominated by shear failure. Reportedly, many mining-
ated with the slow hardening of the concrete. All ray- induced seismic events of large moment magnitude also
paths through the concrete bulkhead, when averaged, have high Es/EP ratios, and can be analysed realistically
suggested 0.18 m/s and 0.13 m/s per day increases in Vp by shear models.
Excavation disturbed zones and their seismic properties 131

shows the recorded ratios of Es/EP and their location in


relation to the advancing face of this carefully excavated
test tunnel. In the region where X/2R  2, i.e. within
two diameters of the advancing face, where the 3D
stress distribution is changing to a 2D distribution,
there are greatest numbers of likely tensile-dominated
events, and a small number with such high S-wave/
P-wave energy ratios (Es/EP), that shearing events may
be suspected. Martin et al., 1997 reported that the
‘slabbing’ associated with the notch formation (Figure
7.22a), started at 0.14 to 0.28  1 diameter from the
face (X/2R  0.14 to 0.28).
Earlier during the URL work, Maxwell and Young,
1996, had reported an interesting case of ‘passive-source’
(i.e. AE) tomography from South Africa. Concurrence of
AE events with high stress conditions ahead of the mining
face in the South African Blyvoor gold mine were again
associated with high velocities (for example 5.8–5.9 km/s)
(a) in the P-wave tomogram. Therefore using passive source
(AE) tomography, the velocity image could potentially
be used to map problem areas. The majority of small
magnitude rock bursts in the mine were located in
regions of high velocity gradient, between a low-velocity
failed zone and a high-velocity, highly-stressed zone.
Logic would perhaps indicate that this was a region of
high shear stresses.

7.3.3 Äspö – EDZ studies

The Äspö hard rock laboratory (HRL) is the second


major location in Sweden for investigating nuclear
waste disposal problems such as the excavation distur-
(b) bance zone. (The first was at the disused Stripa mine,
where many international teams cooperated in SKB’s
Figure 7.22 a) Some 800 microseismic events recorded in the 3.5 m
facility.) At the 420 m level in Äspö HRL, extensive
thick slice that was line-drilled in the floor of the URL
mine-by tunnel. The notch formation is also shown,
seismic and radar investigations were performed, prin-
together with a 2D calculation of the deviatoric ( 1– 3) cipally in order to compare the depth of excavation
stress contours from an isotropic elastic model. b) damage zones in immediately adjacent drill-and-blast
Microseismic event location relative to the advancing and TBM sections of tunnel. The site of the ZEDEX
face of the test tunnel, and the relevant ratio of S-wave (zone of excavation disturbance experiment) is illus-
and P-wave energies (Es/EP). The 78% of ratios 10 trated schematically in Figure 7.23.
suggested dominance of tensile cracking sources. Cai Some of the results of the radial cross-hole tomogra-
et al., 1998. phy performed by Cosma and Enescu, 1996a,b, are
shown in Figure 7.23. The borehole radar and seismic
cross-hole tomography produced comparable locations
Their URL data for the 800 or so events clustered for some major joints, which correlated with core log-
within the line-drilled 3.5 m thick slice, showed that ging in the relatively good quality granite and diorite
Es/EP ratios were most frequently between about 6 and (rock mass quality Q-value mostly 10 to 40, weighted
12, and 78% of events had ratios 10. Figure 7.22b mean  23).
132 Rock quality, seismic velocity, attenuation and anisotropy

There was a very mild EDZ effect on Vp and Vs due


to the high stresses and the partly discontinuous joint-
ing. P-wave velocities measured across horizontal
planes were 6.06  0.1 km/s for both the TBM and
drill-and-blast tunnels. In the vertical planes, differences
in the two cases were observed, 5.96  0.15 km/s around
the TBM drift and 6.26  0.1 km/s around the drill-
and-blast drift. Prior to excavation, the three-dimensional
principal P-wave velocities were 6.06 (vertical), 5.99
(horizontal, NW) and 5.90 (horizontal, NE) km/s
respectively.
Principal stresses at the ZEDEX site were approxi-
mately 32, 17 and 10 MPa. Emsley et al., 1996, showed
that Vp (fast) was parallel to 3 (vertical), while Vp (slow)
was perpendicular to 1 due to anisotropy caused by
jointing. In this case Vp (intermediate) was parallel to
1. In fact, both jointing and the stress-anisotropy
caused the moderate anisotropy in seismic velocities at
Äspö.
In the ZEDEX damage zone studies at Äspö, Bauer
et al., 1996, also recorded the changes of velocity in 3 m
long radial boreholes drilled in the wall and floor of the
TBM and drill-and-blasted drifts, using sonic logging.
Reductions in Vp from the background value of 5.9 to
6.0 km/s were recorded in the first 0.25 m into the
TBM tunnel walls, and up to 1 metre into the walls of
the drill-and-blasted tunnels.
It is of interest to note the ‘flat’ response seen in
Figure 7.24, where tangential stress changes around the
(a)
tunnel are not ‘seen’ by the seismic velocity, due pre-
sumably to the good quality of the rock (Q-value 
22–24) and the existing high stresses, which were high
enough to have acoustically closed the joints, but not
high enough to cause excavation induced micro-cracking,
as shown earlier at the URL (Homand-Etienne and
Sebaibi, 1996).
The rock quality Q c-Vp-M inter-relationships repro-
duced in Figure 7.25 can be utilised to demonstrate some
concurrence with the above data. Some 800 metres of drill
core logged by NGI at the ZEDEX site showed a
weighted mean rock quality Q-value of 24 (‘good’ qual-
ity). If we utilise the c/100 normalisation of Q (Chapter
5) using c  200 MPa as measured, Q c equals approxi-
mately 50. Vp is predicted to range from about 5.8 to
6.2 km/s for an equivalent depth range (relevant to the
principal stress range) of about 400 to 1200 metres, rather
(b) close to the measured values reported above.
Figure 7.23 The Äspö ZEDEX site for TBM and D  B evaluation The intact rock laboratory E-modulus was 69 GPa,
of EDZ effects, with cross-hole seismic tomography and a simple UDEC model showed concurrence with
results. Cosma and Enescu, 1996a,b. the small measured deformations of 1 to 2 mm when a
Excavation disturbed zones and their seismic properties 133

velocity range of 5.8 to 6.2 km/s. Calculated dynamic


moduli around the drill-and-blast drift ranged from
76 to 79 GPa. Presumably the situation M  Edyn
has some implied relation to ‘acoustically closed’
jointing.

7.3.4 Stripa – effects of heating in


the EDZ of a rock mass

The EDZ studies performed in relation to nuclear


waste isolation problems focuses much attention on the
(a)
disturbed properties in this zone. We have seen many
examples of increased disturbance from the frequently
measured reductions in seismic velocity and deform-
ation modulus, and these reductions, taken in conjunc-
tion with increased joint void ratios, generally lead us to
expect enhanced permeability in the EDZ. This does
not always occur however.
In the Stripa SCV (Site Characterisation and Valid-
ation) experiment (Olsson et al., 1993), inflow to the
test tunnel was less than almost all the hydraulics mod-
ellers had predicted, based on dedicated borehole per-
meability testing, using holes drilled along the future
periphery of the tunnel.
Several potential reasons for this discrepancy were
(b) described by the researchers involved, including dissolved
air coming out of solution at the reduced pressures, blast
gas invasion of the joints, and Poisson expansion in the
third (axial) dimension causing increased axial (i.e. nor-
mal) stress on dominant joints crossing the tunnel. Rock
mechanics modelling with two- and three-dimensional
distinct element codes (UDEC-BB and 3DEC) had
shown insufficient shear (mostly 1 mm) for dilation-
enhanced permeability changes, due to the small size
(2  3 m) of the test tunnel, which had quite discontinu-
ous jointing. This property was modelled with ‘numeri-
cally glued’ joint ends. (Barton et al., 1992b).
The planned use of the geosphere as a potential dis-
posal volume for nuclear waste has meant that the heat-
ing (and cooling) effect in the rock exposed in the floor
(c)
or walls of test tunnels has been the subject of much
Figure 7.24 EDZ measurements of Vp in 3 m long radial holes in research. Large diameter disposal boreholes for high level
the TBM section of the Äspö ZEDEX project. Bauer waste canisters will also be in the EDZ of the excav-
et al., 1996. ations, and will create their own smaller EDZ around the
large boreholes. What effects can we expect on local rock
deformation modulus of 60 GPa was used. The equa- stresses, on seismic velocities, and on permeability, due to
tion relating mean deformation modulus M with Vp the production of considerable heat over long time-spans
shown in Figure 7.25 (see inset), suggests deformation in the early disposal period, followed by the cooling
moduli ranging from 58 to 79 GPa, for the predicted period?
134 Rock quality, seismic velocity, attenuation and anisotropy

Figure 7.25 Rock quality Q-value observations from 800 m of core logging, corrected to Qc by the high value of c  200 MPa, gives a real-
istic prediction of Vp and deformation modulus at 450 m depth, when compared with ZEDEX data.

A drift in the Stripa mine in Sweden, used for bore- in velocity at the end of the test (M7–M9). However, it
hole heater tests, showed some interesting effects of might be related to the improved closure of joints at
thermally induced strains. These caused increased seis- elevated temperature, even under constant normal stress,
mic velocities between drained monitoring holes in the as measured by Hardin et al., 1981, and discussed in
jointed quartz monzonite. A schematic of the experi- more detail by Barton et al., 1985.
mental set-up shown in Figure 7.26a, indicates the rela- Upon cooling, the less rough, interlocked joints may
tive locations of the ultrasonic monitoring holes, which ‘spring-open’ more than their closed neighbours, to avoid
were drilled to 10 metres depth, twice as deep as the tensile stress development. This could then cause a
heater borehole. marked reduction in seismic velocity if the open joint or
Paulsson and King, 1980, showed the increases of seis- joints, happened to cross the path of the seismic array.
mic velocity for cross-hole measurements (M8–M6), as a Related local increases in joint conductivity, and
function of time after heater turn-on and turn-off. Pre- reduced shear strength, of any ‘open’ and more planar
and post-heater velocities were generally similar, as shown joints, could be an added uncertainty in nuclear waste
in Figure 7.26b, and a velocity anomaly at about 3 m related disposal scenarios, as emphasised by Barton and
depth was ‘smoothed-out’ by the heating but ‘returned’ Makurat, 2006.
when the rock was cooled. Further details of this Stripa heater experiment
Across the mid-plane of the heater, at 4.2 m depth, the (Paulsson and King, 1980) were subsequently reported in
four sets of cross-hole velocities (Figure 7.26c) showed a comprehensive analysis by Paulsson et al., 1985. The
more or less parallel behaviour, except for an extra strong full duration of the test was 750 days with 398 days of
reduction in velocity for M7–M9 after heater turn-off. heating. The long period of cooling generally returned
The authors give interesting similar-trending curves for seismic velocities to values lower than before the heating,
monitored displacement and stress change about 1 m suggesting permanent changes (such as local excessive
from the heater (Figure 7.26d). They do not have a joint opening as hypothesised above). A significant quan-
confirmed explanation for the anomalous net reduction tity of water expelled during the heating signified a
Excavation disturbed zones and their seismic properties 135

(a) (b)

Figure 7.26 Heater experiments at Stripa, with velocity changes matched by stress and displacement records. Paulsson and King, 1980.

general closing of the joints. Temperatures were over The full record of P-wave and S-wave velocities over
100°C in a small region around the heater and water the 750 days duration of the test is shown in Figure 7.27a
was expelled also from distant boreholes where perhaps and b. The largest velocity changes caused by the heating,
the low initial permeability was less reduced. amounting to 0.2–0.3 km/s were interpreted as occurring
The initial increase in velocity with temperature was in the direction of the minimum horizontal stress, which
linear and varied from 2 to 4 m/s/°C. The average joint is logical since the calculated thermal stress was as much
frequency in the test area, analysed from 224 m of core, as 55 MPa in, presumably, the direction of maximum
was 8.3/m, and an elastic continuum analysis conducted horizontal stress.
prior to the test had indicated larger stresses and local The effect of the heating in an in situ experiment
displacements than were actually measured, presumably such as that described by Paulsson and King, 1980, and
due to the compliance of all these joints. This effect was Paulsson et al., 1985, is to change both the stresses in
also presumably experienced in a ‘heated-mine-by’ the rock and the degree of saturation, particularly close
experiment in the Climax Mine, in the USA. to the source of heat. In an effort to understand and
136 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(a)

(b)

(b)
(c)
Figure 7.27 Complete P- and S-wave velocity record for one of
the Stripa heater tests. Maximum Vp increase was par-
allel to minimum stress. Paulsson et al., 1985.

isolate some of these effects, Paulsson, 1984, conducted


uniaxial tests (0–52 MPa) on granitic (quartz monzonite)
Figure 7.28 Uniaxial tests on intact but micro-cracked Stripa granite
samples from the same Stripa test drift. The samples were
show the important influence of stress and degree of sat-
intact but because of micro-cracks showed strong effects
uration on Vp, and to a lesser extent, Vs. Paulsson, 1984.
of saturation levels on P-wave velocities.
Figure 7.28a shows the effects of axial stress on axially
measured P and S-wave velocities in the dry and saturated 7.4 Acoustic detection of stress
states. The standard deviation of results from eleven speci- effects around boreholes
mens is shown by the length of the crosses.
A clearer indication of the important effect of degree Plona et al., 1997, utilised the effects of compressive
of saturation on Vp at different stress levels is given in stress concentrations around boreholes to investigate if
Figure 7.28b. Lesser effects are seen on Vs in Figure 7.28c. principal stress orientations could be determined by
This particular sample showed stronger effects than the acoustic measurements from within boreholes, at stress
general trend given in Figure 7.28a, and may presumably levels lower than needed for break-out. They referred to
have been due to increased porosity from low aspect- the triaxial tests on sandstone reported by Sammonds
ratio micro-cracks, since the initial velocity when dry et al., 1989, reproduced in Figure 7.29, to emphasise
was only 4.6 km/s (at 5 MPa), and rose to more than the possibility of differentiating the high tangential
5.8 km/s with full saturation. stress from the low tangential stress region around a
Excavation disturbed zones and their seismic properties 137

Figure 7.31 High frequency, axial P-wave monitoring around a


Figure 7.29 High pressure triaxial tests on sandstones showing 10 cm borehole in a uniaxially loaded cube of sand-
the strong coupling of stress with velocity changes. stone. Plona et al., 1997.
Sammonds et al., 1989.

Figure 7.32 Acousto-elastic calculation of Vp anisotropy around


a uniaxially loaded borehole in sandstone ( H 
10 MPa) Plona et al., 1997.
Figure 7.30 Theoretical elastic stress distribution around a uni-
axially loaded borehole. Jaeger and Cook, 1977. Break-out reportedly started at the 15 MPa stress level
where velocity maxima were registered across a diameter.
borehole. An elastic model of the latter is shown in The decline of velocity seen at 19 MPa was due to
Figure 7.30 for the case of uniaxial loading (Jaeger and mechanical damage in the same diametrically opposite
Cook, 1977). max locations. In Figure 7.32, the authors show the
Plona et al., 1997, used a 50 cm cube of sandstone with results of an acousto-elastic model for a borehole in sand-
a central 10 cm diameter borehole loaded uniaxially stone, loaded with a boundary stress of 10 MPa. The
to 21 MPa, to investigate the potential of axial acoustic general similarity of model and experiment is striking.
refraction monitoring at numerous azimuth locations One may wonder whether these effects are taken into
around the borehole. Their principal results are shown account in the general interpretation of sonic logging
in Figure 7.31. down boreholes, since several hundred m/s variations in
138 Rock quality, seismic velocity, attenuation and anisotropy

velocity are seen in a ‘simple’ unjointed sandstone, with anisotropy, plus structure in the form of vertical or sub-
admittedly, an extreme induced tangential stress vertical jointing that would be variously opened or
anisotropy of 30 MPa for max, and (-) 10 MPa for closed by the respective effects of min and max. Is it
min (based on the isotropic elastic tangental stress possible that the sonic logging ‘misses’ such effects of
‘rule-of-thumb’ 3A-B, 3B-A). Significant azimuthal velocity anisotropy around the horizontal plane, due to
velocity anisotropy effects would seem to be possible principally vertically and steeply inclined ray paths?
in wells where there was significant horizontal stress
8 Seismic measurements for
tunnelling

Tunnel face seismic tomography using a pair of bore- meters ahead of an advancing tunnel face. Here, both
holes that was performed by NGI at the Oslo Fjellinjen sources and receivers are placed within the tunnel. By
Tunnel, in the late 1980’s, was used to help the con- the nature of the reflection result, it is difficult to deter-
tractor plan for ground freezing in one tube, while careful mine if the quality will be better or worse. Others have
multiple drift excavation was performed in the other proposed, and demonstrated, the advantages of combin-
tube. The 2  13 m span motorway tunnel beneath the ing this with in-tunnel seismic refraction, with an in-
Oslo downtown harbour district, passed through a wide tunnel source and both internal and external (mountain
regional fault zone with 2 to 4 m of crushed alum shale side) receivers. One can then obtain reliable velocity and
over the arch, underlying 20–30 m of soft, sensitive rock quality predictions ahead of the face, both in front
marine clays. The advantages of the ahead-of-the-event of, and behind the reflectors, which will thereby correl-
data far outweighed the relatively short stoppages and ate better with possible pending tunnelling difficulties.
modest cost involved.
There are now several companies around the world
who are offering the use of reflection techniques for
8.1 Examples of seismic
mapping marked reflectors out to many hundreds of
applications in tunnels

Concerning the obvious need to have information


‘ahead-of-the-event’ in a tunnel, Nord et al., 1991, com-
mented that the present ‘lack of information has only
been accepted due to the high cost of obtaining it’. The
authors went on to analyse the duration and frequency
(in metres) of down-time at some hard rock projects.
They concluded that probe drilling to 50 metres, and 1/2
to 1 hour for the seismic probing measurements would
likely be sufficient. It does not take long to percussion-
drill to 50 m with modern hydraulic jumbos, when 1 to
4 m per minute rates of penetration are achieved.
We will now go backwards in time to an impressive
early example of the use of geophysical surveys in tun-
nels, as given by Scott et al., 1968, for the Straight
Creek pilot bore of 4.0 m diameter, driven under the
continental divide in Colorado, USA in 1963 and
1964. (Figure 8.2). Both seismic and resistivity meas-
urements were made at regular and irregular intervals
along the bore, in order to sample each class of rock.
Five rock classes were defined, based principally on
joint spacing (3 cm to 0.9 m), mineral alteration
(%) and presence of fault gouge, foliation or schistocity.
Figure 8.1 Cross-hole seismic tomography performed by NGI at The rock types themselves (granite, diorite, gneiss,
the face of the Oslo Tunnel in 1987, when approaching migmatite and schist) did not appear to determine rock
a major fault zone. Nord et al., 1991. class. The seismic velocity of the deep layer beneath the
140 Rock quality, seismic velocity, attenuation and anisotropy

Figure 8.2 Cross-section of Straight Creek pilot bore showing geophysical test locations. Scott et al., 1968.

excavation-disturbed low-velocity layer (caused by stress is itself partially correlated with the deep layer velocity
redistribution and gravity induced loosening in the arch) (Figure 8.4b).
was correlated to a number of key construction parame- Already at this early date (the mid 1960s), the authors
ters by the authors, and good correlation was indicated in envisaged a time when geophysical probing ahead of
many cases. The following deep layer velocities were meas- tunnels would have enough correlations to conditions in
ured at the five seismic spreads shown on Figure 8.2: #1 tunnels of different size, and ‘in all environments’, that a
5.2 km/s, #2 5.1 km/s, #3 4.8–6.1 km/s, #4 4.2 km/s and full diameter bore could be driven directly, without the
#5 6.0 km/s. Later, the much lower shallow layer velocities need for time-consuming pilot bores.
will be introduced into the discussion. It is of interest to speculate that the great difficulties
Figure 8.3a and b compare the thickness of the low- encountered when driving the full-scale, twin bore
velocity layer and the so-called ‘tension arch’, defined Straight Creek tunnel were caused by, among other fac-
by the authors as the depth where no further measured tors, an unexpectedly large scale effect caused by the 4 m
dilation of the rock in the arch could be detected. The to 12 m tunnel size difference, and perhaps insufficient
range of seismic velocities shown (14000 ft/s  4.3 km/s, appreciation of the effect of stress on the seismic velocities
20,000 ft/s  6.1 km/s) correspond to the range of rock at that time. There is also the possibility of adverse inter-
classes 5 to 1, and obviously reflect the confinement effect action between the twin tubes, with ‘plastic zone’ overlap,
from the mostly 200–500 m of overburden. a problem of relevance when assessing risk in twin-bore
The worst quality (class 5) corresponded to seismic TBM tunnelling, where conditions are very unfavourable
spread #4 in Figure 8.2 (Vp  4.2 km/s), which had a for any type of tunnelling. (Barton, 2004c).
local overburden of about 300 metres, and obviously The attempted use of steel sets in fault zones at Straight
was strongly affected by the depth or stress level. Our Creek, probably allowed too much (scale dependent)
earlier hard rock relation Vp  3.5  log10 Q with depth loosening of the ground, with the low velocity layers
correction, (Figure 5.36), suggests a possible Qc – value even more affected than in the pilot bore of 4 m span.
of about 0.1 assuming low porosity, hard rock is involved. The seismic velocities of the shallow layers at the five
At the surface a Vp value of 2.5 km/s is implied for this seismic spreads listed above (#1 to #5, Figure 8.2)
rock mass quality i.e. severely faulted ground which were respectively: 3.0 km/s, 2.3–2.7 km/s, 2.3–3.1 km/s,
appears to correspond to the authors’ description of class 1.3–1.6 km/s (worst case, class 5) and 2.3 km/s.
5 rock. (A disturbed zone of at least one diameter was Perhaps more attention should have been paid to these
indicated for such rock.) extremely low EDZ velocities (and to the thicknesses of
Scott et al., 1968, showed good correlations between such zones in the worst rock) which caused almost insur-
Vp and construction data such as steel set spacing mountable problems in the main bore, which took sev-
(Figure 8.3b), and rate of construction (Figure 8.3d). eral year to complete. The Q-system support pressure
They also showed a significant correlation of tunnel sup- database (Barton et al., 1974) includes Straight Creek
port type with electrical resistivity (Figure 8.4a) which main bore as almost the highest recorded tunnel support
Seismic measurements for tunnelling 141

Figure 8.3 Disturbed zone thickness and some support and construction rate details based on Vp measurement at the Straight Creek pilot
bore. Scott et al., 1968.

capacity of at least 300 tnf/m2, and an estimated rock A/3  4.0 to 4.6 km/s, E/3  2.6 km/s. Based on the
mass ‘quality’ Q-value as low as 0.001. Multiple perim- classification of rock conditions (Classes 1 to 6), he gave
eter drifts filled with concrete of some 2 to 3 m thickness tunnel support loads in the range 0.01 to 0.3 MPa (1 to
were needed in some zones. 30 tons/m2 range for Classes 1 to 6), and spacing of the
Ikeda, 1970, assembled a comprehensive set of tech- steel arch support in the range 1.5 to 0.75 m, and con-
nical data from some 70 rail tunnels in Japan. He first crete thickness in the range 0.3 to 0.9 m, for 30 m2 and
classified rock types into classes A to F, as reproduced in 60 m2 tunnel sections from his 70 case records. These
Table 8.1. Examples of Class A rocks were metamorphic data are reproduced in Figure 8.5 a, b, c and d. This is
rocks such as gneiss, quartzite, etc., while examples of a valuable set of early case records and their technical
Class E were Pleistocene rocks such as mudstone and description, using seismic velocities.
volcanic ejecta. A somewhat finer division of rock types than the
He then listed typical ranges of seismic velocities (Vp) original Japanese Railways classification of Ikeda, 1970
under rock conditions (Classes 1 to 7), shown in Table 8.2 has been used in more recent years by the Japanese
for each of the previous rock type classes A to F. The Highway Authority. This is reproduced in Figure 8.6
two examples under Class 3 rock condition would be (from Barton and Itoh, 1995), showing the addition of
142 Rock quality, seismic velocity, attenuation and anisotropy

Figure 8.4 Some cross-correlation between support type, resistivity and undisturbed (deep layer) velocity. Straight Creek pilot bore. Scott
et al., 1968.

Table 8.1 Rock type classes (A to F) of Ikeda, 1970 based on lithology and geology.

Rock qualities Names or rocks

a) Paleozoic rocks Mesozoic rocks clayslate, sandstone, greywacke, conglomerate, chert, limestone, ‘schalestein’, etc.
b) Plutonic rocks granite, granodiorite, diorite, gabbro, peridotite, etc.

A c) Hypabyssal rocks granite porphyry, quartz porphyry, porphyrite, diabase, etc.


d) Volcanic rocks (some part) rhyolite and andesite of Mesozoic era, basalt, etc.

e) Metamorphic rocks gneiss, hornfels, schist, phyllite, quartzite, etc.

a) Metamorphic rocks having


conspicuous schistocity
B
b) Paleozoic and Mesozoic rocks
having fine bedding planes

a) Mesozoic rocks (some part) shale, clayslate, sandstone, tuff breccia, ‘schalestein’, etc.

C b) Palaeogene rocks (some part) silicified shale and sandstone, tuff breccia, welded tuff, etc.
c) Volcanic rocks (greater part) rhyolite, dacite, andesite, basalt, dolerite, etc.

a) Tertiary rocks mudstone, shale, sandstone, conglomerate, tuff, tuff breccia, welded tuff,
D agglomerate, etc.
a) Pleistocene rocks Neogene rocks mudstone, siltstone, sandstone, sand and gravel rock, tuff, terrace, talus, fan,
E volcanic ejecta, agglomerates, etc.

F a) Alluvium rocks Diluvium rocks clay, silt, sand, sand and gravel, loam, volcanic ejecta, fan, talus, terrace, etc.
Seismic measurements for tunnelling 143

Table 8.2 Rock condition classification using seismic P-wave velocities based on 70 Japanese rail tunnel case
records analysed by Ikeda, 1970. (Note writer’s interpretation of some of author’s words under
‘Remarks’ and ‘Notes’).

Rock qualities
Classification of
rock conditions A B C D E F

1 5.0 4.8 4.2


2 5.0–4.4 4.8–4.2 4.2–3.6

good
3 4.6–4.0 4.8–4.2 4.4–3.8 3.8–3.2 2.6

medium
4 4.2–3.6 4.4–3.8 4.0–3.4 3.4–2.8 2.6–2.0
5 3.8–3.2 4.0–3.4 3.6–3.0 3.0–2.4 2.2–1.6 1.8–1.2
6 3.4 3.6 3.2 2.6 1.8 1.4–0.8

bad
7 1.4 1.0

Remarks:
1) Always go to a poorer class when ground water will inflow into tunnel.
2) Rocks with expansive nature have no relation to this classification.
3) Raise 1 or 2 classes when the Poisson’s ratio of the weathered rock is better than 0.3.
Notes:
1) The numbers 1–7 are the rock condition ratings.
2) The numerals show the velocities of elastic wave in the rock (km/s).
3) See Table 8.1 for the rock qualities A through F.

(a) (b)

(c) (d)

Figure 8.5 Relationships between support intervals (steel sets), concrete lining thickness and support pressure, each as a function of rock
condition classes 1 to 7, which were defined by Vp ranges. Ikeda, 1970.
144 Rock quality, seismic velocity, attenuation and anisotropy

a preliminary rock quality Q-scale at the base of the fig-


ure. A comparison of the velocity ranges shown in
Figure 8.6, with Table 8.2 shows, of course, broad simi-
larity in view of the common, relatively young geology,
that has to be tunnelled by the two state authorities.
Note that the appended Q-scale is based on the sim-
ple shallow, hard rock equation (Vp  3.5  log Q),
with no immediate attempt to correct for uniaxial
strength through Qc, nor to adjust for significant matrix
porosities. The ‘fit’ of Vp with Q will therefore be
increasingly in error in the case of the lower velocity,
weaker, and more porous rocks, which require individ-
ual interpretation of these (usually) inter-related factors,
which collectively have a strong Vp-reducing effect.
This means that a low velocity Tertiary sandstone
with a shallow-seismic refraction velocity of 2.5 km/s,
may have an initially implied Qc value of only 0.1,
while the implied Qc value is 1, if n  15%. (Figure
5.36). A significantly higher apparent Q-value (due to
c  100 MPa) is implied, concerning the usual fac-
tors describing the structure of the rock mass, like
RQD, Jn (number of joint sets), Jr and Ja. However if
the rock is incompetently weak, RQD is zero (accord-
ing to the definition of Deere et al., 1967), but a min-
imum of 10 in the Q-calculation to avoid an impossible
Q  0. Furthermore, and most importantly, there may
be an unfavourably large ratio of stress-to-strength
in the context of tunnelling, giving the need for an
elevated SRF (see Appendix A).
The differences seen above are due to a fundamental
difference between characterization (with no excavation
involved), and a classification, which has become known
in tunnelling, as a scheme for selecting appropriate rock
support and reinforcement, via RMR (Bieniawski,
1989), or Q.
An account of pre-investigations and experiences
while driving Norway’s first sub-sea road tunnel to the
west-coast island of Vardø, was given by Palmström, Figure 8.6 The Japanese Highway rock-mass-class and rock-type
1982. Seismic refraction profiles totalling almost 13 km tabulation, together with the anticipated seismic P-wave
covered a 500 m wide zone across the 1.5 km wide velocities from shallow refraction seismic. Approximate
sound and gave depth to bedrock with an accuracy rock mass quality Q-scale from Barton and Itoh, 1995.
of 0.5 m in relation to 36 soundings in the sea bed.
The presumed and actual weakness zones encountered in Figure 8.7. The overall distribution of Vp is also shown
during driving of the 9.4 m span tunnel are shown in in Figure 8.8, for a rock mass with a mean logged
Figure 8.7. The rock cover (mean 40 to 50 m) and shallow Q-value, outside the weakness zones, of the following:
sea depth (20–40 m) are also shown. Up to 4 joint sets
90 1 1
were logged in the quartzitic sandstones, siltstones and Q     1.9
shales, which had frequent clay coatings on the generally 12 4 1
steeply dipping joints, giving poor stability, especially in Palmström, 1982, recorded a 560 m length of concrete
the low velocity (e.g. 3.2–3.3 km/s) crushed zones shown lining (350 m placed at the face), 2500 m3 of shotcrete
Seismic measurements for tunnelling 145

Figure 8.7 Norway’s first sub-sea tunnel to Vardø, showing assumed and encountered weakness zones. Palmström, 1982.

40

35

30
Distribution

25

20

15

10

0.14
3000 4000 5000 6000
(a) Seismic rock mass velocity

(b)

Figure 8.8 P-wave velocity statistics for Vardø sub-sea tunnel. Lower velocity zones and corresponding support methods. Palmström, 1982.
146 Rock quality, seismic velocity, attenuation and anisotropy

Figure 8.9 Vp-Qc-M correlations with depth and porosity correction. Barton, 1995.

and 168,000 rock bolts giving an average tunnelling


progress of 17 metres/week and a cost of USD 8000 per
metre. The weighted mean value of Vp outside the weak-
ness zones was close to 5.0 km/s which, according to
Figure 8.9 (Barton, 1995) implies a Q-value, perhaps
equal to 15 if we assume 50 m of cover and a nominal
1% porosity for these generally hard rocks.
This estimate of rock mass quality from Vp compares
poorly to the quoted value of about 2, which might of
course be in error. Another reason for this discrepancy
might be stress-related, or an effect of the hard beds of
quartzitic sandstones, masking the lower velocity,
weaker beds of siltstones and shales.
The low-velocity zones (Vp  3.2–3.3 km/s) also cre-
ated greater tunnelling difficulties than the values would
suggest, implying an artificially elevated velocity in rela- Figure 8.10 Support costs (as % of total cost) versus Vp, for some
tion to rock quality. This will be discussed further. road, pipeline and water tunnels. Nilsen, 1998.
There is obviously a broad correlation between tunnel
support costs (as a percentage of total costs) and seismic Errors in interpretation of seismic measurements, for
velocity, as shown for example by Nilsen, 1998, for the example due to horizontally interlayered stiff (meta-sand-
case of half a dozen sub-sea tunnels (Figure 8.10). Support stones) and softer shales, have on occasion caused unwel-
costs may rise from 50% to at least 75% of total costs, come surprises, i.e., with ‘false’ velocities, apparently as
when the P-wave velocity reduces from 5.5 to 4.5 km/s. high as 4.5 to 5.0 km/s, nevertheless requiring immediate
As pointed out by Nilsen, the most important factor is the cast concrete lining up to the tunnel face. In one such case,
quality of pre-investigation and follow-up, and an under- the depth effect upon Vp shown in Figure 8.9, may have
standing of the need for good investigations by the owner. been responsible for some of the ‘false’ velocity increase in
Seismic measurements for tunnelling 147

Figure 8.11 Rock cover in relation to seismic velocities at sub-sea


tunnel sites. Nilsen, 1998.

a generally extremely dry sub-sea tunnel (Barton and


Monsen, NGI contract report concerning North Cape
Tunnel, 1997, for Norwegian Road Authority).
The logged rock mass quality Q-values of only 0.01
and the need for heavy support right up to the face
were ‘inconsistent’ with the general range of Vp  Figure 8.12 Mean P-wave velocities at some tunnel and cavern
4.5–5.0 km/s. In this case the rock load and the sea sites in Norway, with shotcrete and concrete lining
depth may each have contributed, due to the undrained frequencies. Sjøgren et al., 1979.
situation.
Nilsen, 1998, also gave rock cover versus bedrock 15 to 55% concrete lining and from 20 to 50% shot-
depth (water  sediment) data for numerous sub-sea crete lining (with rock bolts).
tunnels together with associated P-wave velocities. As The mean velocities given in Figure 8.12 are of
expected, the highest velocities were usually associated course a bit misleading, as it is the lower values in any
with the lowest rock covers, while the zones with lowest given distribution of velocities that require the rock
velocities (as low as 2 km/s) tended to have higher support. For example the few shotcrete lined sections of
designed rock cover, as shown in Figure 8.11. the Mongstad oil storage caverns (case 1, Figure 8.12)
Sjøgren et al., 1979, performed shallow refraction would certainly have had a lower, local velocity than the
seismic investigations at sites where subsequent sub- mean value of 6.0 km/s for these massive, foliated meta-
surface excavations were planned or actually carried anorthosites.
out. They were therefore able to show several cases As the authors point out using an illustrative statistic
where seismic velocity showed a broad correlation with from the Vardeåsen site in Norway, the high velocities
tunnel support measures such as shotcrete and concrete (4.8 to 6.2 km/s) completely dominate the usual range of
lining. velocities from these relatively unweathered Scandinavian
Although excavation span is listed in Figure 8.12, the sites, and it is the much smaller number of tectonic zones
authors did not give depth, for the caverns or tunnels (shear zones, faults), dykes and joint swarms with veloci-
listed. Significant depth is probably the reason why ties from about 2.5 to 3.5 km/s that cause the construction
velocities between 3.8 and 4.7 km/s (which might corres- problems, especially when high inflows of water occur.
pond to hard rock Q-values of about 2 to 15 at 25 m If we utilise the shallow-depth, hard rock, low porosity
depth) apparently were associated with such major Q-Vp conversion given below:
stability problems at Rendalen (cases 5 to 8), where dif-
ferent sections of the headrace tunnel required from Q  10( Vp 3.5) (8.1)
148 Rock quality, seismic velocity, attenuation and anisotropy

Figure 8.13 Geological profile and rock classification details for Tunnel A. Mitani et al., 1987.

the above velocity ranges for good rock and fault or frac- less than the nominal 100 MPa for hard rocks. We can
ture zones can be approximately converted as follows: therefore use the porosity (n%) and Qc  Q  c/100
corrections in the Qc–Vp chart, which then help to
1) Vp  4.8 to 6.2 km/s (Q  20–500)
explain the somewhat higher Q-values.
2) Vp  2.5 to 3.5 km/s (Q  0.1–1)
Cecil, 1971, warned of another source for the possible
As can be seen from Figure 8.9, a velocity as high as lack of correlation between seismic velocity and rock
6.2 km/s may for reasons of depth of measurement (or quality, when considering the presence of thin clay coat-
unusually high rock strength) only imply a Q-value of say ings in otherwise widely spaced jointing. The latter cause
25, so conversions using this equation must always be potential stability problems but may hardly change
related to the more comprehensive Q–Vp–depth–porosity RQD or Vp values. On the other hand, the Jr/Ja terms in
model reproduced in Figure 8.9. the Q-system may capture the correct stability problems
At the Veas sewage treatment plant near Oslo, Norway, by reducing the Q-value, but Q will then not correlate so
where seismic surveys were performed by Geoteam well with the velocities. These potential pitfalls in the
(Sjøgren, 1984), and logging of excavations in the nodu- general, near-surface correlation for hard rocks (equation
lar limestone were performed by NGI colleagues, the 8.1) need to be carefully evaluated from case to case.
reported correlations between Vp and rock quality Q were
as follows:
8.2 Examples of the use of seismic
Vp (km/s) Rock quality Q data in TBM excavations
3.0 1 to 5 (with support)
4.4 20 (no support needed) A good example of the application of seismic velocity
measurement to interpret TBM penetration rates is given
For hard rock at shallow depth, with negligible poros- by Mitani et al., 1987. They investigated the rate of
ity (central trend, Figure 8.9), the above velocities would advance of two small diameter TBM tunnels (  2.6 m)
have suggested somewhat lower Q-values than the above by measuring Vp and Schmidt hammer rebound values
(0.3 and 8 respectively). However, the nodular limestone, for the wide variety of rocks encountered. Tunnel A
consisting of inter-bedded and well cemented shale and (shown in Figure 8.13) was mostly driven in sandstones,
limestone layers, has some porosity (5 to 10% could be slates, porphyry and weathered conglomerates, with
estimated) and its uniaxial compressive strength ( c) is generally strongly developed jointing and poor stability.
Seismic measurements for tunnelling 149

(a)

(a)

(b)

Figure 8.14 a) Vp (solid bars), and Schmidt hammer rebound %,


in two Japanese TBM tunnels of 2.6 m diameter,
b) correlation of Vp to steel rib support spacing. Mitani
et al., 1987.

Tunnel B was driven in more homogeneous but variously


weathered granites.
Figure 8.14a shows Vp in relation to Schmidt hammer
rebound %, while Figure 8.14b shows the degree of rock (b)
support in relation to the Vp values. (Rock support con-
sisted of unsupported sections, or various grades of steel Figure 8.15 Net penetration rates as a function of Schmidt ham-
mer rebound and seismic velocity for 2.6 m diameter
rib, steel channel or I-beam, with successively reducing
TBM tunnels in Japan. Mitani et al., 1987.
spacing, i.e., 1.5, 1.2, 1.0, 0.8 m as Vp reduced.
Penetration rates, ranging from 0.6 to 5 metres per hour
are shown correlated with seismic velocities in Figure 8.15. As can be noted by the above, the range of assumed
Utilisation of geomechanics and seismic testing rock qualities (approx. Q  0.1 to 15) do not penalise
to correlate with TBM penetration rates (PR  advance rates by any time-consuming support needs, so
net-instantaneous, and AR  effective advance rate), was effective rates of advance for this hydropower tunnel
described by Sampaola et al., 1978. Figure 8.16 shows a were inversely ‘correlated’ to two of the geomechanics
quite sensitive correlation between progress in m/hr and measures given above, and were therefore effectively
rock mass class A, B or C (which represented statistically inversely correlated with Q-values. This is consistent
homogeneous zones). The tunnel was only 6.4 m in diam- with the QTBM model of Barton, 2000.
eter, and was bored in granites of variable quality caused
by alteration and variations in jointing frequency. The 8.3 Implications of inverse
depth range for the tunnel was not given by the authors. correlation between TBM
The TBM appears to have been a little under-powered in advance rate and Vp
relation to the strongest, least jointed rock mass class.
The set of data given by Sampaolo et al., 1978, can In view of the intended aim of correlating, where pos-
be reproduced approximately as shown in Table 8.3. sible, the seismic velocity and the rock quality, it is
150 Rock quality, seismic velocity, attenuation and anisotropy

appropriate to consider in more detail the inverse cor- of the figure. Faster boring will correlate with lower Vp
relations seen in the two previous case records. We have values up until some limit, as suggested by the two
seen from Figure 8.15 the strong correlation between descending portions of the drilling rate trends.
support needs and seismic velocities, which follow the Until the above support/stability limit is reached, the
normal trend of increased support with lower Vp-values, net penetration rates (PR) seen in the comprehensive data
and with lower assumed Q-values. of Mitani et al., 1987, may be considered to have the
However, it will be noted that net penetration rates approximate upper and lower bound values given in Table
(PR), correlate inversely with Vp values, in other words, 8.4, in relation to Vp and assumed, shallow depth (nomi-
the increased degrees of jointing and reduced strength nal 25 m), hard rock Q-values.
(also seen in the Schmidt hammer results) help to increase
the penetration rate. The same trend is seen in Table 8.3
for a larger TBM (  6.4 m) boring in granites.
The documented trends of degree of jointing and
rock strength on drilling or boring rate seen in the
above examples are summarised in diagrammatic form
in Figure 8.17. The same type of inverse correlation
with Vp values can be envisaged by converting the rock
mass quality Q-class to a Vp-class in the upper portion

Figure 8.16 Correlations between rock mass class A, B, C and excav-


ation speed for a TBM driven hydropower tunnel.
Respective classes had Vp  5 km/s, 3.5–4.5 km/s, Figure 8.17 Conceptual inverse correlation of boring or drilling
3.5 km/s, and c 150 MPa, 50 MPa and 8 MPa. rate with Q-value or Vp-value. Modified from Barton,
Sampaolo et al., 1978. 1996b.

Table 8.3 Correlations between advance rates and seismic velocities (Sampaola et al., 1978), with last column added by writer, using central
trend for 100 m depth, or 50 m depth, from Figure 8.9.

Effective advance Net rates Qest.


Zone Alteration Vp (km/s) c (MPa) rate (m/hr) (AR) m/hr (PR) F m1 (Barton, 1995)

A. Sound granite little or no alteration 5.0 150 0.4–0.6 1.0 2 8 or 15


B. Jointed granite medium degree of alteration 3.5–4.5 50 1.8–2.0 2.5–5 2, 5 0.07–2, 0.2–4
C. ‘Cataclastic rock’ high or very high alteration 3.5 8 2.5–3.0 4–5 5 0.07, 0.2
Seismic measurements for tunnelling 151

Table 8.4 Upper and lower-bound PR–Vp–Q trends from Mitani


et al., 1987 with additional interpretation.

Net penetration rate (PR)

Upper-bound Lower-bound

PR (m/hr) 1.0 2.0 3.0 4.0 1.0 2.0 3.0 5.0


Vp (km/s) 6.0 5.0 4.0 3.0 4.5 3.0 2.0 1.5
Q 300 30 3 0.3 10 0.3 0.03 0.01

8.4 Use of probe drilling and


seismic or sonic logging ahead Figure 8.18 Conceptual drawing of sonic probe and conversion of
of TBM tunnels data to rock mass Q-values and rock support classes.
Barton, 1996b.
The last decade of developments with double-shield
TBM’s that use the PC-element liner for thrust when linkage complex in many cases, the approximate expect-
resetting the grippers, have made it possible to tunnel ation that final displacement (, millimetres) is approxi-
through a range of geological conditions, generally with mately given by:
less delays. However, the complexity of the machines
has drawbacks in commissioning and learning-curve
time, and a simpler design could in principle be used, if SPAN(m )
 (8.2)
prior information of ground conditions were available Q
through probe drilling and geophysical measurements.
The concrete element liner, while convenient from may be an invaluable correction to the Vp-Q correlation,
many points of view, may be an expensive solution if which may have a ‘stress problem’, concealing the actual
most of the rock is actually of very good quality, requir- poor quality. (See Barton, 2000, for detailed correlations
ing only light support. between , tunnel dimension, Q-value and other stress-
Nord et al., 1991, indicated that TBM advance rate related factors).
could be optimised in mixed ground conditions, by Unfortunately, the fact that the rock mass Q-value
always selecting the right tunnelling mode in response logged in a TBM tunnel may be higher than that
to advance information. The other needs for advance logged in a drill-and-blasted tunnel will affect the meas-
information relate of course to support needs and to the ured deformation as described above. (Q will appear to
possibilities of water inrush or caving ahead of the face. be higher and  will be smaller). Likewise, if refraction
There are numerous TBM around the world that are seismic measurements were performed along the wall of
stopped for long periods (some few, even permanently), the TBM tunnel, the values of Vp obtained would also
due to inadequate information about pending adverse tend to be higher than in the equivalent drill-and-
rock conditions. (Barton, 2000) blasted tunnel for at least two reasons:
In Figure 8.18, the concept of advance seismic velocity
information for subsequent rock quality class and tunnel 1) Reduced level and depth of damage in wall of
support class estimation is presented. Ideally this should TBM tunnel
be made a routine operation, made by fast percussion 2) Higher tangential stresses closer to the tunnel wall.
drilling, such that support components such as steel
arches, rock bolts or shotcrete can be immediately avail- Such aspects will influence other details of the behav-
able, and applied with appropriate timing , at and behind iour of the rock mass, due to coupled behaviour. For
the tunnel face. example there will be a tendency for lower permeability
A more complete concept is illustrated in Figure 8.19. and less drainage around the TBM tunnel, which, for rea-
Here, displacement monitoring is also performed in an sons of more complete saturation might also increase the
effort to roughly confirm the Vp–Q correlation. Since seismic velocity. However, a seismic velocity probe ahead
special depth or anisotropic stress effects make the Q-Vp of the tunnel will not see the difference between the TBM
152 Rock quality, seismic velocity, attenuation and anisotropy

Figure 8.19 Conceptual use of sonic probe for rock mass Q-value estimation, with displacement monitoring for confirmation of support
needs. Barton, 1996b.

tunnel and the drill-and-blast tunnel, unless the measure- conditions, appear to have taken a lead in the use
ment was made too close to the face of the tunnel. of seismic for ‘probing ahead’ of their tunnels, in par-
ticular TBM tunnels, where the consequences of
8.5 In-tunnel seismic measurements delays are more critical, due to both the investment
for looking ahead of the face level and the normal expectation of fast tunnelling.
Following on from Mitani et al., 1987 analyses of
The Japanese, with an extremely active tunnelling the relation between TBM progress and seismic veloc-
industry, in combination with far from ideal geological ity, reviewed earlier, we will briefly refer to some
Seismic measurements for tunnelling 153

Figure 8.21 Seismic refraction principle, for both in-tunnel and


surface sources, with both in-tunnel and surface
Figure 8.20 Seismic reflection for identifying a change of conditions, receivers. Hayashi and Saito, 2001.
combined with the results of sonic logging in a probe
hole. From Kajima Corp. Nishioka and Aoki, 1998.

The authors suggested that this situation could be


more recent advances concerning seismic ‘probing rectified by using the in-tunnel refraction estimates of
ahead’. velocity distributions ahead of the face, so that the
The advantage of being fore-warned of changed con- accuracy of reflector positions could be increased. With
ditions by means of in-tunnel reflection methods is the necessary velocity distribution ahead of the face, the
nicely illustrated in Figure 8.20. Also shown is the use rock mass could be characterized both up to and
of a probe hole, and a sonic logging of this hole, with a beyond the better-located reflectors.
low velocity fractured zone correlating with an interval The authors justified their method by demonstrating
of reduced ‘breaking energy’, possibly actually referring the steadily improving match to a hypothetical moun-
to compressive strength, in view of the units used. tain velocity model, as numbers of in-tunnel sources
Clearly, the reflector further ahead, presents the possi- were increased. The model, and two stages of improve-
bility of either reduced or increased rock mass quality. ment, are reproduced in Figure 8.22.Theoretical travel
A radically reduced Vp and Q-value and a decision times were calculated by ray-tracing, and were con-
for pre-injection, pre-reinforcement and perhaps over- sidered like observed data.
boring may save weeks or months of delay and cost only The authors applied their proposed method to a tun-
one or two days in ‘lost’ production. Such an investment nel under construction, in Mesozoic slates, sandstones
appears worthwhile. and chert, with an overburden varying from 100 to
Hayashi and Saito, 2001 described an interesting 300 m. Figure 8.23 shows the detailed surface refrac-
approach to seismic surveys for tunnelling that is a tion seismic model of the mountain terrain, which was
logical extension of conventional high-resolution sur- produced before tunnel construction. A general veloc-
face refraction seismic, namely the use of sources and ity along the tunnel route of about 4.0 to 4.2 km/c was
receivers also at, and close behind, the tunnel face. The indicated at this stage. The subsequently installed
concept is shown schematically in Figure 8.21. GPS in-tunnel sources and receiver are also indicated, together
clocks are needed to synchronise the sources within the with the string of surface receivers down the ‘opposite’
tunnel and the receivers at the surface. mountain face.
The authors pointed out that the already developed An in-tunnel reflector method that was being used in
reflection method of HSP or TSP (horizontal or tunnel this tunnel, had imaged a clear reflector some distance
seismic profiling, e.g. Sattel et al., 1992), with both ahead of the face when the face was at 439 m. The in-
source and receiver in the tunnel, locates seismic reflec- tunnel source refraction method subsequently utilised
tors ahead of the tunnel face, as desired, but that the with only one in-tunnel source, predicted a sharply
reflector distributions are not related to rock quality declining velocity ahead of the face. In fact a face col-
directly. It is also difficult to determine if the rock qual- lapse occurred at chainage 544 m, 105 m ahead of the
ity will get better or worse at a given reflector, and they measurement location, and a 300 m wide zone of weak
suggest that there may be inaccuracies of location in rock, with velocity as low as 3.7–3.8 km/s was indicated
view of the unknown actual velocity field. by the second method.
154 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(b)

(c)

Figure 8.22 a) A velocity model with ideal number of in-tunnel sources and limited surface sources. b) Two sources within the tunnel por-
tals, and limited surface sources, improves the reconstructed velocity model, with improved definition and location of the low
velocity zones. c) A series of mostly systematic tunnel sources, but with a central gap, gives greatly improved match to the
‘actual’ case. Hayashi and Saito, 2001.

8.6 The possible consequences of pressures, at depths of 700–900 m. TBM tunnel


insufficient seismic investigation (diameter 5 m) eventually ran sub-parallel to individual
due to depth limitations faults, causing delays of at least half a year for each
1 m wide fault (AR  0.005 m/hr). TBM finally
In the final section of this chapter a TBM case record will abandoned; new contractor for D  B from other end
be briefly referred, in which the depth of tunnel (mostly of tunnel.’
700 to 900 m), and mountain-side screes and loose The tunnel was plagued by these sub-parallel valley-
deposits, had hindered the correct interpretation of the side faults for at least 2 years, with up to 6 months fault-
actually very adverse structural conditions. The case also related delays on several occasions, until TBM tunnelling
illustrates the problem with stress effects, or virtual ‘seis- was abandoned. An attempt to detect the continued pres-
mic closure’, on this occasion due to compaction and ence of a particular fault (see sketches from daily
near-invisibility of actually very troublesome faults. geological log in Figure 8.25) using seismic tomography
Figure 8.24 indicates the initial geometric difficulties of between two divergent pilot boreholes, proved to be
this valley-parallel, TBM-driven headrace tunnel, for the unsuccessful, due to the presumed confining and densi-
Pont Ventoux hydroelectric project in north-west Italy. fying effect of the high stresses (from 800 m of over-
This drawing of a valley-parallel fault swarm was devel- burden) on the fault-zone materials.
oped several years after the original investigation, and In Figure 8.26, the rock quality Qc-based velocity-
shows the limited surface seismic that was attempted at the depth model is shown again, this time with some
time, together with some insufficiently deep boreholes. appended comments concerning the possibly elevated
The case was summarized in Barton, 2004c as velocities of highly confined fault zones. Such zones,
follows: ‘Unpredicted fault swarm parallel to valley- despite their Qc values as low as 0.01 or even 0.001, can
side, together with very high (and fault-eroding) water nevertheless exhibit a stiffness and compactness at
Seismic measurements for tunnelling 155

(a)

(b)

Figure 8.23 a) Pre-construction surface seismic result, showing the in-tunnel sources for in-tunnel and surface receivers. b) An in-tunnel
reflector method had indicated a reflector ahead of the face and a small reduction in velocity was assumed. The in-tunnel source
refraction method subsequently predicted a sharply declining velocity ahead of the face, and a face collapse occurred about
105 m ahead of this location. Hayashi and Saito, 2001.

Figure 8.24 Original seismic refraction profiles and inadequate borehole depths, are compared with the geologist’s later re-assessment of the
actual valley-parallel fault swarm, that had a dramatic effect on the fate of the TBM, and the final decision ‘to drill-and-blast’
from the other end of the tunnel. Pont Ventoux, Italy.
156 Rock quality, seismic velocity, attenuation and anisotropy

Figure 8.25 Plan and elevation views of the ‘2750’ fault at Pont Ventoux, based on a super-position of the site geologists sketches of the
developing situation. This was a case of an unexpected combination of high stress, high permeability, and high fault frequency,
and the eroding power of high pressure water. Barton, 2004c. Attempts to ‘probe ahead’ using seismic tomography between two
diverging boreholes, proved not to be as successful as expected, due to the relative ‘invisibility’ of the assumed extension of the fault.

depth, that makes them nearly ‘invisible’. But in fact 2.3–2.7 km/s, 2.3–3.1 km/s, 1.3–1.6 km/s (worst case,
there is hope that they will still show a recognisable class 5) and 2.3 km/s.
contrast to the even higher velocities of surrounding The worst quality (class 5) corresponded to seismic
rock, as suggested in the labels in this figure. spread #4 in Figure 8.2 (deep layer Vp  4.2 km/s),
There is an interesting support for the above logic which had a local overburden of about 300 metres, and
from the first case record referred to in this chapter, obviously was strongly affected by the depth or stress
namely the continental divide Straight Creek pilot tun- level. At the surface a Vp value of 2.5 km/s is implied for
nel in Colorado, described by Scott et al., 1964. It may this rock mass quality i.e. severely faulted ground which
be recalled that the in-tunnel seismic refraction had appears to correspond to the authors’ description of
been differentiated into ‘deep layer’ and ‘shallow layer’ class 5 rock. (A disturbed zone of at least one diameter
velocities. was indicated for such rock.).
The following deep layer velocities were measured at the So in fact one may conclude that the deep layer veloc-
five seismic spreads shown on Figure 8.2: #1 5.2 km/s, ities bore no resemblance to the eventual major tun-
#2 5.1 km/s, #3 4.8–6.1 km/s, #4 4.2 km/s and #5 nelling difficulties experienced when excavating the 12 m
6.0 km/s. The seismic velocities of the shallow layers (i.e. span twin tunnels. The deep layer velocities were either
the loosened, near-surface-of-the-tunnel layers) at the five undisturbed (but highly stressed), or were perhaps sub-
seismic spreads listed above, were respectively: 3.0 km/s, ject to additional tangential stress compaction effects.
Seismic measurements for tunnelling 157

Figure 8.26 Rock mass quality Qc – Vp – depth model, showing the potentially elevated P-wave velocities of nevertheless seriously-delaying
fault zones, if encountered at great depth, as at the Pont Ventoux head-race tunnel. Contrast to the even more elevated velocity
of the surrounding ‘country rock’ can nevertheless be expected. Barton, 2004c.

The more relevant, extremely low EDZ velocities depth or stress effect that masks, in velocity terms, the
from the pilot bore investigations, truly representing true low quality. The 300 m overburden at this collapse
the poorer rock classes, , were actually what caused the location, would from Figure 8.26, suggest a near-surface
almost insurmountable problems in the main bore, Vp of about 2.5 km/s – i.e. relevant to a serious fault
which took several year to complete. Multiple perim- zone, or extremely poor rock.
eter drifts filled with concrete, making some 2 to 3 m Finally, one may note the adverse effects of low
effective wall thickness were needed in some zones. Q-values on TBM progress, shown in Figure 8.27,
One may also note from the Hayashi and Saito, 2001 specifically because of fault-zones. Velocity measure-
case record, reproduced in Figure 8.23, that the face ments at depth may not suggest such low values of Q.
collapse at 544 m chainage, actually occurred in a The TBM may nevertheless be delayed.
Vp  4.1 km/s rock mass. This is ‘illogical’, without the
158 Rock quality, seismic velocity, attenuation and anisotropy

Figure 8.27 The typical performance trends derived from analysis of 140 TBM tunnels, with ‘unexpected events’ strongly tied to low rock
mass quality Q-values. Barton, 2000. It is probable that extremely low actual Q-values might show a deceivingly ‘high’ range of
P-wave velocities, in the case of imaging ahead of deep tunnels.
9 Relationships between Vp,
Lugeon value, permeability and
grouting in jointed rock

Since matrix porosity, and in particular joint porosity, each data of Saito, 1981 (Figure 2.17) and by the use of
affect the permeability of a rock mass it is perhaps logical the time average equation for dry and saturated chalk
that the seismic P-wave velocity should show some degrees (Grainger et al., 1973). The approximate ‘porosity’ which
of correlation with the permeability. When the joint void appears in the time average equation may contain air or
space is artificially increased by a particular type of ‘per- water, and this porosity obviously affects the overall veloc-
meability’ testing (i.e., high pressure Lugeon injection ity, i.e., whether Vair  0.33 km/s or Vwater  1.44 km/s
tests), stronger correlation with velocity can be expected, is involved. So we have a theoretical starting point for a
due to the lowering of effective stress. This may apply most saturated velocity. The key question is whether this helps
strongly in the case of rock masses of poor quality that in predicting possible permeabilities. Does the saturated
are easily deformed around the injection boreholes, during velocity give any indication of actual flow resistance?
the Lugeon tests, but the possible presence of clay is a com- Extensive sets of in situ measurements of rock foun-
plicating factor. The unit Lugeon is defined as the number dation moduli, permeability and seismic velocity were
of litres per minute flowing from each metre of a double- assembled by the Comité National Français, 1964,
packered section of borehole, under an excess injection from numerous dam site investigations. For the special
pressure, above the groundwater pressure, of 1 MPa. case of two sites in jointed granite (from France’s Massif
Since a rock mass generally contains joints and micro- Central), a strong correlation was evident between Vp
cracks, which are both a source of water and compliant, and the Lugeon test results.
the application or existence of an anisotropic (effective) Figure 9.1 shows an approximately linear distribution
stress distribution may preferentially have closed those of data on a semi-log plot of Vp versus the Lugeon value.
oriented at an obtuse angle to the major stress, while
keeping those at acute angles or sub-parallel to the major
stress, ‘open’. There are then grounds for expecting both
anisotropic permeability tensors, and a corresponding
anisotropic velocity, with both maxima tending to be
parallel or sub-parallel to the major stress. Since sub-
vertical jointing may dominate in the same way that hor-
izontal stress anisotropy may dominate, the anisotropy
will tend to be related to azimuth. However, there is a
potential source of error here. The permeability test holes
must intersect the ‘open’ structure to register their higher
permeability. The test holes needs to be drilled in the
‘slow’ direction, parallel to the minimum stress. The
lower Q-value given by crossing all the ‘open’ joints,
should correspond to the higher permeability.

9.1 Correlation between Vp and


Lugeon value Figure 9.1 Evidence for a correlation between Vp and Lugeon
value at two granitic dam foundations (Comité National
In Chapter 2, the strong effect of saturation on P-wave Français, 1964) (Q-value scales have been added by the
velocity was convincingly demonstrated by the extensive writer.)
160 Rock quality, seismic velocity, attenuation and anisotropy

If we make the assumption that shallow refraction seis-


mic, or relatively shallow cross-hole measurements of
velocity were used, then we can tentatively investigate the
relation Vp  3.5  log10 Qc (the diagonal line in Figure
9.1) as a means of relating Q-value and Lugeon value.
In very approximate terms we can see from the data
that not only proportionality, (i.e., L  1/Q) but equal-
ity L  1/Q is evident in the approximate range 100 to
1.0 Lugeon. The scatter of velocities and Q-values is
seen to be about one order of magnitude, in other
words neither velocity measurement, not Q-logging
must be substituted for the testing. However, L  1/Q
(Lugeon) might be utilised in extrapolation exercises,
or to identify non-conforming behaviour.
We may therefore tentatively write:
1
L(Lugeons)  (9.1)
Q

as a useful approximate for fitting to data in some rock


masses, and for explaining deviation (i.e. channelling,
in other cases). There is an upper-bound Vp-L trend in
the French data that exactly parallels the Q  1/L
envelope; as can be clearly seen in Figure 9.1.
Two campaigns of core drilling at a shallow tunnel Figure 9.2 Correlation between Vp and Lugeon values at four
site in Wales; first with vertical holes, then with 45° hard rock sites in Norway. After Sjøgren et al., 1979.
(Q-values scale added by the writer. Stippled curve
inclined holes to intersect more of the steep structure,
given by equation 9.1).
gave mean Lugeon results of 12 and 28. If we assume
Q  1/L, then Q values of 0.08 and 0.036 are derived.
Completely independent Q-logging of the relevant which suggests a Lugeon value of 8 when Vp is 2.6 km/s.
boreholes (8 from phase 1 vertical holes, and 13 from The measured value in the relevant holes was 12.
phase 2 inclined holes) by the writer, gave weighted The above logging data from clay-bearing metasedi-
mean Q-values from many hundreds of observations of ments shows remarkable similarities to the L  1/Q
the six Q-parameters that were as follows: model, and also shows the potential anisotropy of the
Q-value due to different joint sampling frequency with
Q (BH 1 to 8)  0.11 (‘higher’ Q, lower permeability:
hole orientation. Lower Q-values, higher Lugeon values,
L  12) and lower seismic velocities will tend to be measured
Q (BH 13 to 21)  0.08 (‘lower’ Q, higher permeabil- when perpendicular to major structure. The opposite
occurs when paralleling major structure. Of course there
ity: L  28)
can be exceptions to this basic concept caused primarily
The tunnel itself showed an overall weighted mean Q- by an eventual rotated stress-anisotropy, that no longer
value of 0.05, i.e. it was rather unstable rock. Downhole matches the joint patterns: a less likely scenario.
Vp logging in BH 1 to 8 gave a mean Vp  2.6 km/s for Sjøgren et al., 1979, gave correlations between Lugeon
the same depth range that was core logged. This con- tests and seismic velocities for several locations from
verts to a predicted Q-value of 0.12, almost the same as four of their investigated hard rock sites in Norway. A
logged. By using the following ‘hard rock’ method of total of 29 data points are given in Figure 9.2. They
conversion, based on Vp  3.5  log10 Q and L  1/Q, defined 1 Lugeon in the usual way, and mentioned the
and eliminating Q we obtain: constant pressure of 1 MPa. It is not known if this stand-
ard excess pressure was reduced closer to the surface,
L  10(3.5 Vp ) (9.2) but if not, this could be the reason for some of the
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 161

Table 9.1 Approximate correlations between measured transmissivities and seismic parameters based on measurements
in five boreholes in marl. Albert, 2000.

Transmissivity [m2/s] 1014 1012 1010 108 106 104 102

Vp [m/s] 5500 5117 4733 4350 3967 3583 3200


Vs [m/s] 3000 2700 2400 2100 1800 1500 1200
Vp/Vs 1.72 1.84 1.96 2.09 2.21 2.33 2.45
Dynamic shear modulus [GPa] 24 20 16 12 7 3 0.1
Dynamic E-modulus [GPa] 60 50 40 30 20 10 0
Poisson’s ratio 0.28 0.30 0.32 0.35 0.37 0.39 0.41

unexpectedly high values (e.g., 10 to 20 Lugeons) seen scales) with the logarithmic transmissivity scale. (The
in rock with velocities from 3.5 to 4.5 km/s. However, transmissivity is the product of the permeability and
in general their results showed the expected lower the thickness of the measured aquifer or ‘aquiclude’.)
Lugeon results at higher velocities, as in the French data. Most of the measurements reported by Albert, 2000,
The rule of thumb 1/Q  number of Lugeons was were in the transmissivity range 105 to 1012 m2/s. A
tested against the Sjøgren et al., 1979, data using the 1 m long test section assumption would convert these
assumed, near-surface, hard rock relationship Vp  to the more familiar engineer’s m/s units.The borehole
3.5  log10 Q (Barton, 1995). It appears from Figure depths ranged from about 400 m to 1800 m, and
9.2 that the ‘1/Q’ curve is a suitable lower bound to included faulted and brecciated rock.
some of the velocity-Lugeon data. Accuracy could poten- One must assume that the ‘hydraulic tests and fluid
tially be improved if Lugeon values were correlated with logging’ were of the non-deforming type, unlike the
depth zones, and if these depth zones had been given, a civil engineering Lugeon testing (often for evaluating
more correct Vp–Q relationship could have been groutablity), discussed elsewhere in this chapter. Com-
selected from Figure 8.9 (see Chapter 8). paction effects on this relatively weak rock at borehole
Curiously, some of the points plotted by Sjøgren et al., depths up to 1800 m presumably have affected perme-
1979, exactly fit the dotted line (Q  1/L) relationship, abilities more than the seismic velocities. The smaller
as also experienced in Figure 9.1, possibly indicating steps in velocity at the lower transmissivities resemble
conforming or non-conforming data, as the case may be. the effect of depth in the Qc-M-Vp engineering model
In poorer quality rock masses, a Lugeon test is a form (Figure 8.9). The successive reductions in dynamic
of rock mass deformability test, with unusually sensi- E-modulus with increased transmissivities (and Qc-
tive (‘aperture cubed’) registration of joint deformation, values?) have a certain similarity to this rock engineering
due to the tendency, under laminar flow conditions, for model, bearing in mind the Edyn  M inequality.
flow rates to be proportional to e3, where e is the Examples of correlations between seismic refraction
hydraulic aperture. However, in very stiff rock masses surveys and drilling and tunnelling results are given by
this may not apply, especially if channel flow dominates Sjøgren, 1984. This example is given in this chapter
due to outwash channels in filled joints, or if joint jack- due to permeability links. Figure 9.3 shows successive
ing (slight opening) occurs due to low Ko ( v/ h) val- stages of an investigation, and confirmation during con-
ues, or if jacking occurs due to over-dimensioned water struction, for a water supply tunnel beneath the Skien
pressures close to the surface. river in Norway. Four seismic refraction profiles are shown
From extensive work in marl formations in in the top figure. Three low velocity zones were indicated
Switzerland, Albert, 2000, indicated quite strong rela- beneath the river, the largest of which (Vp  2.5 km/s)
tions between selected seismic parameters and transmis- was proved by an inclined borehole to be a partly con-
sivity measurements in five deep boreholes at Wellenberg, solidated breccia and loose alum shale (core loss aver-
a potential nuclear waste repository site. Good correla- aged 75% in this zone).
tions with transmissivity were obtained with Vs, dynamic The Lugeon value in this zone was 14, which might
shear (
) and E-moduli, Vp, Vp/Vs and dynamic correspond to a Q c value of about 1/14 (0.07). This
Poisson’s ratio. is close to the value of Q that could be predicted from
Table 9.1 shows the author’s approximate correl- Figure 8.9, using a nominal porosity for the zone of
ations between the seismic parameters (using linear 5%, and the 50 m depth shown in Figure 9.3c. At this
162 Rock quality, seismic velocity, attenuation and anisotropy

is related to national or regional practice.) The number


of Lugeons is expressed by the well-known relation L 
litres/min/m/1 MPa excess pressure. Most of the flow
losses (and joint deformation) occur close to the bore-
hole in such a test, which differs greatly from the care-
ful, low pressure pumping (or extraction) tests favoured
in permeability testing (Quadros, 1995).
By good fortune or correct physics, the modulus of
deformation (M) that was shown in Figure 8.9 is pro-
portional to Q1/3 or to Q 1/3
c in the case of rocks weaker or
stronger than our nominal c  100 MPa. Similarly, it is
well known that flow rate is more or less proportional to
e3 in jointed rock masses (where e  equivalent hydraulic
aperture of the joints, and intrinsic permeability can be
expressed as e2/12). The smaller value of (e) approaches
the physical aperture (E) when e  1.0 mm, and this
inequality (E/e  1) is related to joint roughness JRC
(Barton et al., 1985, Barton and Quadros, 1997).
Around the injection borehole we may assume that
the natural joint apertures are deformed significantly,
especially when maximum injection pressures of 0.025
up to 0.1 MPa per metre depth are used. The latter
European injection pressure limit at dam sites is about
two times the assumed vertical total stress. When ko
( h/ v) is 1.0 and causes lower minimum stress
than these figures, some slight hydraulic jacking of
some of the joints is an obvious consequence in the ini-
tial radii around the boreholes.
Figure 9.3 Correlations between refraction seismic velocities and
The following basic assumptions will be made con-
borehole and tunnelling experiences through the same
cerning this all-important joint deformation region
zones. Sjøgren, 1984.
around the injection holes:
depth (and with n  5%) Q  0.07 corresponds to
1. The Lugeon value (L) which is recorded as volumet-
Vp  2.5 km/s, as measured by chance or good physics.
ric flow rate (litres/min) will tend to be proportional
The tunnel was driven through the same zone, in the
to the cube of the new apertures that have been cre-
direction of profile 1 (Figure 9.3c). Grouting was neces-
ated, i.e., ( E3). There is some evidence from
sary for the Vp  3.0 km/s zone (perhaps Q  0.3 and
grouting results (over-coring or excavation) that the
L  3 from equations 9.1 and 9.2). Probe drilling and
most permeable and well-connected joints open
heavy reinforcement was used through the 12 m wide
most at the expense of others in the same set. The
fractured alum shale zone, which had Vp  2.5 km/s
resulting Lugeon value will often be dominated by
(and L  14, and Q  0.07?).
the Emax value and we can roughly approximate here
3
that L  Emax , since the smallest micron-size aper-
9.2 Rock mass deformability and tures will contribute only minutely.
the Vp-L-Q correlation 2. The locally gapped joint will have an aperture Emax
that is approximately inversely proportional to defor-
Dam sites throughout the world are investigated by mation modulus M.
means of borehole water injection tests, typically using 3. The calculation of a ‘double’ Boussinesque elastic
double packers, and injection-pressures related to foundation calculation for the radially distributed
depth below the surface, but usually limited to about deformation of each side of the joint, with realistic
0.25 or 0.5 or 1.0 km/cm2 per metre depth. (The choice input for dimensions, supports this.
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 163

Therefore we have the following possible inter-rela- the chalk marl and quoted permeabilities from Lugeon
tionships between maximum apertures, Lugeon values, type tests of 104 to 106 m/s in these areas. If we
deformation moduli and Q-values, which in turn are assume for simplicity that 1 Lugeon  107 m/s then
linked to seismic velocities: the very high Lugeon values obtained of 1000 to 10
imply Q c values of 0.001 to 0.1 according to equation
L  E3max 9.1 (where Q has been replaced by Q c).
Emax  M1 These low Q c values can be converted to ‘tunnel sup-
M  Q1/3
c port’ Q values of 0.02 to 2 if we assume a mean c value
Therefore L  Q1
c of 5 MPa for the chalk marl. This range is in line with
expectations for the heavily jointed rock mass at
(Note ‘’ implies ‘approximately proportional to’ in the Chinnor. (The term ‘tunnel support Q-value’ is used to
above proportionalities). remind of the original development of the Q-system for
These simple proportionalities therefore suggest that selecting tunnel support: Barton et al., 1974.)
the number of Lugeons may indeed be proportional to Although one should in general resist the temptation to
1/Q c, unless other mechanisms than local joint defor- convert 1 Lugeon to 107 m/s (approx.) as if ‘rock mass
mation are responsible for the flows, for example out- permeability’, because deformability of the medium is
washed chlorite fillings, severely canalised flow due to very likely in the case of Lugeon testing in weaker rocks,
basalt flow-top weathering, uncontrolled hydraulic joint it is nevertheless of interest to note that the 1000 to 0.001
jacking, and so on. The data that follow the lower-bound Lugeon scale shown in Figures 9.4 and 9.5, would con-
trend L  1/Q in Figures 9.1 and 9.2 are therefore con- vert to 104 m/s to 1010 m/s. This resembles the wide
sistent with this theoretical model, and explain why Vp range of permeability often encountered where thousands
and L (Lugeon) can show a degree of correlation. of well tests are assembled in one plot.
In Figure 9.4, the above inter-relationships have been However, channelled flows in weathered basalt flow
expressed in the form of a nomogram, using the basic tops may exceed 102 m/s, and some massive igneous
structure derived in Chapter 5 (Figure 5.36). Five illus- rock may have permeabilities as low as 1012 m/s, due to
trative ‘type curves’ have also been added, to show what lack of joint connectivity and lack of micro-cracks. (The
might be typical Vp-Q-M-L data for fault zones, weak latter may appear only after sampling from strongly
porous rock, hard jointed rock and hard massive rock. anisotropic virgin stress states, as discussed in Chapter 3).
For simplicity (and continuity of the curves), it has been A further example of Q-Vp-L correlation can be
assumed that porosity only develops in the shallow, developed from the columnar basalt foundations of the
near-surface weathered zone in each case (H  25 m). Segunda Angostura dam site in Argentina. Classification
In the case of a hard porous rock, there will be a velocity of the site together with preliminary testing were reported
correction caused by porosity at all depths, and this will by Di Salvo, 1982.
mean that the central curve shown in Figure 9.5 will
give the correct velocity, roughly midway between the mean RMR  63
‘hard rock’ reference curve (top) and the porosity cor- mean Q  8.5
rection curve (bottom).
The steepening gradient of the type curve is in this (These are close to the Barton, 1995 suggested inter-
case affected by the assumed Q-value increase at greater relationship RMR  15 log10 Q  50)
depth. If, for some reason, this does not occur, then the Vp (downhole)  4.5 km/s below ‘decompressed zone’
‘type curve’ could be a straight vertical line, as for the Vp (downhole)  2.0 km/s in the ‘decompressed zone’
‘hard massive’ rock shown in Figure 9.4. Any porosity
correction would merely reduce the ‘height’ (i.e., The higher velocity suggests Q c  10, based on the
reduce the maximum velocity) of such a line. The Qc relation Vp  3.5  log10 Q for hard rock. A uniaxial
correction factor is the same as that developed in strength for the basalt of e.g. 125 MPa, would mean
Chapter 5 (Figure 5.36). The nomogram can be illus- Q  8.
trated by the following ‘coupled’ example. L (Lugeon) in decompressed zone  16, suggesting
At the Chinnor Tunnel in chalk marl, Hudson et al., Q c  0.06 based on L  1/Q c.
1980, (‘seismic’ Hudson) referred to very low velocities A Q c-value of 0.06 suggests a Vp value of 2.3 km/s,
(0.6 to 1.0 km/s) for badly fractured/jointed areas of i.e., very close to the measured velocity.
164 Rock quality, seismic velocity, attenuation and anisotropy

Figure 9.4 Potential inter-relationships between Vp, Q, M and L, with corrections for depth, porosity and compression strength. Barton, 1999.

Figure 9.5 Hard porous rock of 10–20% porosity. Example type curve for estimating Vp-Q-M-L data.
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 165

9.3 Velocity and permeability Another in situ block test, this time in jointed sand-
measurements at in situ block stone in Colorado, USA was reported by Swolfs et al.,
tests 1981. The block was 2 m3 in volume and contained a
near-vertical joint. The joint was calcite filled, and
During the 1970s and 1980s, a series of large scale, in appeared to be about 1.5 mm wide at the surface. The
situ block tests were performed by Pratt and co-workers P-wave velocity of the surrounding jointed rock of
in the USA, in order to evaluate test methods and about 1.5 km/s appeared to be independent of joint fre-
instrumentation suitable for nuclear waste disposal quency and orientation. This is surprising in view of
projects which were being planned at that time. The the presumably drained state of the test site (Figure
block tests were designed to give large-scale properties 9.7a). However, ‘moist’ laboratory samples had about
(1 to 3 metres scale) under controlled loading condi- the same value of Vp.
tions (using flatjacks), and some were at elevated tem- In situ stresses of about 1 MPa were relieved by line
peratures (using borehole heaters). drilling of three sides of the block. The long side of 2.3
The effects of stress application on velocity and per- metres and 1.2 m depth was parallel to the joint. This
meability in jointed granite were first investigated on a resulted in Vp and Vs changing from 1.5 and 0.8 km/s
large scale by Pratt et al., 1977, who used a flat-jack to 0.9 and 0.5 km/s respectively. Calculated values of
loaded block measuring 3  3  3 m, which con- Edynamic thereby changed from 3.3 to 1.2 GPa, assum-
tained three sub-parallel, vertical joints. The rock was ing a rock density of 1.97 gm/cm3 , because the sand-
an anisotropic, but quite massive granite, and the site stone has a high porosity of 25%. The uniaxial strength
was in Wyoming, USA. The authors investigated vel- was about 11 MPa, and static Young’s modulus was
ocity changes as a function of applied stress (0 to 9 MPa) 2.3 GPa, based on laboratory samples. The block was
applied either parallel or perpendicular to the jointing loaded uniaxially (normal to the joint) and biaxially,
(so-called E-W or N-S velocities, respectively). using multiple flatjacks in each of the three slots.
Results for different measurement lengths, including The effect on P-wave and S-wave velocities is shown in
0.15 m long laboratory samples, are shown in Figure 9.6. Figures 9.7c and d. Pre-excavation velocities (shaded
The lab samples, which may have experienced micro- lines) were reached at about 1 MPa. This is exactly the
cracking on release by drilling, show the strongest Vp stress acting when undisturbed velocities were measured.
response to stress increase. Although the 3 m cubed block An anomalous increase in joint deformation was also
was released on all four vertical sides, the ‘contact’ with recorded above this same stress level of about 1 MPa.
‘virgin’ rock stresses along its intact base may presumably The authors also applied shear stresses to the joint by
be the reason for less response of Vp to stress along these activating the flatjacks at the end of the block, while
in situ measurement lengths of 1.0 and 2.85 metres. The holding a constant normal stress across the joint (0.7 or
block also remained nearly saturated, compared to the 1.4 MPa). Since the block was attached at its base, joint
lab sample ‘0.15 m, D’ (D  dry, S  saturated). shearing was limited (even at the top surface of the
More details of the in situ response of rock mass block) to about 0.7 mm, which represents pre-peak
velocity to increasing stress are given in Figure 9.6c. strength. Dilation was negligible (10
m), and is per-
Increased wetting of the surface of the block caused the haps the reason why Vp and Vs slightly increased during
small (0.1 km/s) increases in velocity seen between the application of shear stress to 3.0 MPa, probably mostly
pairs of curves 4 and 9, 1 and 8 respectively. The authors in response to the simultaneous application of normal
finally presented a composite plot (fully coupled behav- stress of 0.7 or 1.4 MPa (Figure 9.7a). If significant
iour) of joint displacement (mm), flow rate along joint dilation had occurred during increased shearing, a
J1 (cm3/s) and velocity, each as a function of stress. reduced velocity would presumably have resulted. The
Figure 9.6d shows that increased closure of the joint small velocity response to moderate stress change seems
after about 2 to 3 MPa normal loading, caused a plateau to be a feature of relatively unjointed, porous rock.
on the permeability-stress curve, and a sharp reduction The authors also performed a permeability test using
in the sensitivity of Vp to further stress increase, espe- injection in a central hole that intersected the joint.
cially beyond 5 MPa loading. This is consistent with They calculated a permeability of 3.7  107 m/s.
stress-Vp data reviewed in Chapter 5, and broadly in There are several interesting coincidental values of
line with the non-linear effect of depth on velocity the reported tests that we can compare with the Qc -Vp-
shown in Figures 5.36, 9.4 and 9.5. M-L model (Figure 9.4). If we follow the ambient
Figure 9.6 Vp changes caused by loading a 3  3  3 m block of granite containing vertical joints, and laboratory tests of the same rock.
a,b) Velocity-stress behaviour for three types of loading conditions, and for three measurement sizes. c) Nomogram linking effects
of uniaxial joint closure stresses with joint J1 deformation D4, velocity across jointed block, and flow rate along part of joint J1.
Pratt et al., 1977. Note tendency for acoustic closure beyond 5 MPa.
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 167

P-wave velocity of 1.5 km/s at the ambient stress of about


1 MPa (equivalent to about 20–25 metres of overbur-
den) in the lower left-hand corner of the Q-Vp-M-L
chart, we find a Q c value of about 0.4 at 25% porosity.
Independent of this, the 1/Q c model for Lugeon estima-
tion suggests a back-calculated Q c-value of 0.27. This is
very close to the velocity-based estimate. The low uni-
axial strength of 11 MPa means that the Q-value can be
estimated as about (0.4 or 0.27) 100/11  3.6 or 2.5.
These are close on a six-order of magnitude Q-scale.
The estimation of deformation modulus (M) can be
based directly on Vp according to Figure 9.4. Thus we
see that 2 GPa is estimated, which is close to the labora-
tory value of 2.3 GPa, and to the Edynamic estimates of
1.2 GPa (unloaded) and 3.3 GPa (undisturbed, loaded
to approximately 1 MPa). In this case this deformation
modulus estimate is based on Vp (Figure 9.4, right-
hand column of M values derived from):

( Vp 0.5 )/ 3
M  10 (9.3)

and this gives a more accurate estimate of 2.1 GPa when


Vp  1.5 km/s. The relevant modulus value is also
obtained using the direct equation between M and Q:

M  10 Q 1/3
c (9.4)

which again gives an estimated 2.1 GPa, when using


Q  0.01. We refer to Q as ‘Q-prime’ since it has not
been corrected for porosity. The real Qc value needs the
porosity correction, and final correction for the ratio
c/100, to reach the assumed rock mass quality Q,
which we estimated from both velocity measurement
and independent Lugeon testing as ranging from about
2.7 to 3.6.
Further checks on rock mass quality can be made the
direct way by using the authors’ descriptions of the
jointing; three sets, spaced at 0.6, 0.9 and 0.3 metres,
with the most prominent set filled with about 3 mm of
calcite. Via the volumetric joint count of Palmström,
1983, we can calculate Jv  6.1, and RQD  95%.

Figure 9.7 a,b) Loaded block test in (drained) unit of in situ sand-
stone containing a vertical joint, loaded on three sides
by flat-jacks. c,d) Vp – and Vs – stress trends for uniax-
ial and biaxial loading, compared with pre-slot veloci-
ties – shaded. e) Effect of joint shearing on Vp at two
different normal stress levels. Swolfs et al., 1981.
168 Rock quality, seismic velocity, attenuation and anisotropy

Figure 9.8 Permeability – stress coupling for three bituminous coals,


due to the detailed cleating or jointing: an extreme ana-
logue for jointed rock masses. Somerton et al., 1975.
Figure 9.9 a) Permeability-Vp coupling for two of the bituminous,
cleated coals. b) Velocity-mean stress coupling for one
The independently estimated Q-value is therefore of the cleated coals. Somerton et al., 1975.
approximately as follows:

which will also be present in jointed rock masses at large


RQD J J
Q   r  w scale, when in situ effective stress states are altered by
Jn Ja SRF large scale pumping or injection experiments.
95  100 1  1.5 1 Three bituminous coals having large differences in
    1.7  2.8
8
9 6 1 hardness and degree of jointing (cleats, etc.) showed
almost equally great sensitivity to applied stress level,
All of the above estimates are very close, considering despite their five order of magnitude range of perme-
the logarithmic (six orders of magnitude) Q-value rock abilities (0.1 to 100 millidarcys). Somerton et al., 1975,
quality scale. We have thus demonstrated that Q, Vp, applied mean stresses over the range 1 to 14 MPa and
M and L are inter-related, and that we may be able to noted between two and three orders of magnitude
include the Lugeon value in this inter-relation, if care is reduction in permeability (Figure 9.8).
taken to eliminate irrelevant non-deforming, channel Simultaneous monitoring of ultrasonic velocity
flow cases. The implication is that depth or stress level, showed increases of velocity of about 0.3 to 0.6 km/s
also an axis in Figure 9.4, also plays an important role (from 1.8 km/s when stress-free) for each order of mag-
in these mutual inter-relationships. nitude reduction in permeability. This is shown in
Using an analogue material for heavily jointed rock, Figure 9.9a together with the Vp-stress behaviour of one
namely coal, one can also see how there is great potential of the coals in Figure 9.9b. Both these figures indicate
sensitivity between velocity, stress level and permeability, greatest changes in Vp and permeability at the lowest
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 169

stress levels (and lowest velocities), just as found in rock minute (10 to 50 microns) shear displacements on a set
masses, due to improved acoustic coupling across joints. of near-vertical joints that were not aligned to H(max).
The joint roughness was assumed to create some increase
in permeability despite the assumed small shear dis-
9.4 Detection of permeable placements.
zones using other geophysical During a stimulation experiment in which the reser-
methods voir was kept ‘inflated’ by a well-head pressure of
6 MPa and a flow rate of 9 litres/s, the velocity showed
Since the mid-eighties, researchers working at nuclear small reductions in the depth zone between 2100 to
waste related rock laboratories such as Stripa, Äspö, and 2500 metres. The seismic data suggested that the per-
Grimsel have utilised both seismic and radar tomogra- manently stimulated cracks and joints were dilating as
phy to characterize major fault zones. Their studies the pore pressure increased, even though the pore pres-
have generally helped to explain why these relatively sure was only 20% of that required for jacking (30 MPa).
small volumes of fractured (or heavily jointed) rock are A certain degree of joint aperture increase can be expected
responsible for such large percentages of the total flow from the elevated pore pressure, whether or not signifi-
of water. At the Grimsel site, Martel and Peterson, cant shearing was occurring.
1991, found that seismic velocity tomography delin- Aoki et al., 1991, describe the use of cross-hole seismic
eated major geologic structures better than radar atten- measurements to compare with (and verify) the direc-
uation or radar slowness tomography. However, they tional distributions of hydraulic diffusivity. In the case of
point out that anomalies on tomograms can reflect a tests in a heavily jointed rhyolite, the lower velocity zones
wide range of features (rock types, alteration, changed (3–4 km/s) between two of the boreholes corresponded
porosity) besides different degrees of jointing or frac- quite closely to the location of highly permeable zones
turing. Knowledge of geological background data is between these boreholes at 6 and 13 metres depth, as
therefore helpful in making better interpretations. seen in tomographic plots of cross-hole test data.
Injection of brine for tracing flow paths has been suc- It is well known that low resistivity measurements
cessful in many projects. Martel and Peterson, 1991, correlate with zones of increased water content and fre-
found that radar attenuation difference tomograms quently with higher permeability. At a site in South
were more reliable in locating brine than slowness Korea, where the writer logged a series of boreholes in
tomograms, at the Grimsel US/BK site. Radar and seis- weathered granites, the opportunity arose to compare
mic signals are sensitive to different physical parameters these independently derived Q-parameter statistics
(mechanical stiffness and electro-magnetic wave con- with resistivity tomograms that were given to the writer
ductivity, respectively). For this reason, the respective after his draft report was delivered.
tomograms highlight different features of the rock It was found that sections of the boreholes with
mass. Radar may delineate permeable zones (porosity increased joint frequency (low RQD, high Jn) did not
caused by pore space or by joint apertures) in slightly always correlate with low resistivity and vice versa, as
different locations to the low seismic velocity zones was reasonably to have been expected. The parameters
associated with clay filled discontinuities. The one will that did show a consistent correlation with low resistiv-
usually lie parallel to the other, since higher permeabil- ity were the low values of Jw (estimated, for example,
ity may be associated with the heavily jointed zones from iron staining or apparent aperture) and the high
that are often found adjacent to faults. This was a phe- values of Ja (for example from sand or silt fillings and
nomenon that at first made geophysics teams question due to clay fillings).
each other’s coordinates at the Stripa SCV (Site The latter gives low resistivity due to the ionic effects
Characterisation and Validation) site (Olsson., 1992). of the clay, since water content (and permeability) are
Green et al., 1989, described the use of vertical seis- clearly lower in such discontinuities than in those that
mic profiles and cross-hole seismic surveys at the are sand or silt filled. There is therefore in fact a poten-
Camborne School of mines 2 km deep hot dry rock geot- tial source of error in judging the meaning of low resist-
hermal project. The objective was to show that the micro- ivity zones.
seismicity generated during stimulation experiments This end of the rock mass quality spectrum is also
corresponded to regions of enhanced permeability. The unfortunately the region where the usual link of low Vp,
majority of micro-seismic events were interpreted as low rock mass quality Q-value and high permeability
170 Rock quality, seismic velocity, attenuation and anisotropy

also may break down, due to the ‘adverse’ effect of clay


on permeability. We will see in Chapter 16 that a par-
tial solution has been found for this clay-based phe-
nomenon, by rearrangement of the function of two
Q-parameters, namely the reversal of Jr with Ja in a sim-
ple term called ‘QH2O’. This depth-dependent model
appears to provide a realistic, preliminary estimate of
permeability to many kilometres depth.

9.5 Monitoring the effects of


grouting with seismic velocity

A simple illustration of the benefits of seismic velocity


monitoring at grout injection jobs was given by By,
1988, using cross-hole (average velocity) measurements
at a dam abutment in Norway. Blast damage from
reflected waves caused by a nearby quarry for rockfill
was suspected to have caused shearing and dilation
along adversely dipping foliation planes, giving Vp
measurements as low as 0.5 km/s in the drained, 5 to 10
metres depth zone.
Following extensive surface injection (Figure 9.10)
Figure 9.10 Superficial and deeper-layer grouting at a dam site
Vp locally increased to between 2 and 6 km/s. However,
abutment, where nearby (and too close) quarry blast-
the depth interval of 7 to 15 metres did not show
ing had caused suspected shearing along the dipping
acceptable velocities (only 2 to 3 km/s) and this was foliation planes, resulting in a (drained) P-wave veloc-
confirmed by additional water leakage tests. A strong ity of only 0.5 km/s. Note the dramatic improve-
depth-velocity effect was observed at the site (velocities ments in the (assumed) foundation properties as a
rising from 2 to 5 km/s from about 10 to 25 m depth) result of grouting. By, 1988.
presumably related both to apparent rock quality
improvement at depth, and to a post-stressing effect
from the increasingly confined grout at greater depth. relation L  1/Q Lugeon. Velocities above 4.0 km/s
Rodrigues et al., 1983, also refer to the correlation (Q  3? or K  0.3 Lugeon?) could not be improved
between seismic velocity and reduced permeability upon by the grouting. Such results emphasise the rea-
achieved by grouting at the Cabril dam site founded on sons for combined use of high injection pressures and
granite, in Portugal. The dam had been grouted 30 micro or ultrafine cements, if e.g. 0.3 Lugeons (or rather
years previously and had already quite a high P-wave 4.0 km/s), should be improved upon. Barton, 2004a.
velocity (range 4.2 to 5.5 km/s), partly as a result of this At the 270 m high Inguri arch dam in Georgia,
earlier foundation treatment. The new round of grout- Savitch et al., 1983, used the seismic velocity criteria
ing increased Vp by 2 to 20% and reduced the permea- shown in Figure 9.11 for judging the success of grout-
bility, as registered by Lugeon testing, by anything from ing. One can first interpret that very high pressures
40 to 100%. A larger grout take (where there was pre- must have been used here, since it is implied that vel-
sumably a larger rock mass ‘porosity’), also corres- ocities as high as 4.5 km/s could be improved by grouting.
ponded to the locations where the largest increases in However, the depth effect on Vp (e.g., Figure 9.4) is
Vp were registered, following the grouting. probably playing a role here. A Vp value of 4.5 km/s
It was noted by Grujíc, 1974, at the 220 m high implies Qc  10 in near-surface, hard un-weathered
Mratinje dam (shown in Figures 4.1 and 4.2) that effect- rocks. However at the 270 m high dam, deep injection
ive consolidation grouting could be performed when Vp grouting and deep Vp monitoring (say at 100 to 200 m
was in the range 2.5 to 3.5 km/s (i.e., approximately depth) might have caused a depth (or stress) related
Q  0.1 to 1.0 or 10 to 1.0 Lugeon according to the enhancement that was equivalent to a much lower rock
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 171

Figure 9.12 Before, and after grouting Vp measurements in sand-


stones and marls, showing increased velocity, and
Figure 9.11 Grouting efficiency (I  excellent, II  good, III  increased anisotropy. Capozza, 1977.
Satisfactory, IV  unsatisfactory) based on velocity
monitoring at the Inguri arch dam. Savich et al., 1983.
earlier. These are hardly values that would justify grout-
ing. Presumably some of the velocity increase caused by
quality of Qc  1, or even less, which was likely to be grouting is due to the increased stress, and some due to
injectable. Savich et al., 1983, results are therefore read- reduced volume of joint apertures and better seismic
ily understandable when the Vp-Q-depth effect is taken coupling. Velocity monitoring alone may therefore not
into account. guarantee a good (i.e., low leakage) grouting result.
Grouting at the Zavoj hydro electric project in Bernabini and Borelli, 1974, describe a variety of
Yugoslavia was monitored by cross-hole velocity meas- early seismic tests performed at hydro electric projects in
urements and by cross-hole seismic tomography per- Italy in the 1950s, 60s and early 70s. In ‘stratified rocks’
formed between two galleries. Slimak et al., 1991, they show about 35% increase in seismic velocity caused
showed three-dimensional contour plots of velocity dis- by successful grouting with cement. However, they
tributions before and after grouting, and a difference observed that the scatter of data did not change; the
tomogram showing the net gain in velocity, as a result stratified rock maintained its anisotropic character, just
of the grouting. The authors emphasised the efficiency the mean velocity was increased (1.96 to 2.65 km/s).
of such measurements in checking the effectiveness of a The before-and-after grouting measurements of Vp
large-scale injection programme. referred to by Bernabini and Borelli, 1974, are reproduced
Unfortunately, the authors did not give the results of in more detail by Capozza, 1977. The inter-bedded sand-
Lugeon testing before and after the grouting. The velocity stones and marls experienced an average velocity increase
increase of only 0.25–0.75 km/s and the relatively high of 0.7 km/s (2.0 to 2.7 km/s) as a result of the grouting.
velocity (mostly 4.0–5.5 km/s) before grouting suggest Since the cross-hole measurements performed before and
only a moderate result. One may speculate that these after grouting gave a range of ray-path angles () in rela-
velocities are also affected by stress level, since ‘near- tion to the gently dipping bedding, it was possible to
surface’ Q-values of 3 to 100 derived from the above vel- show the influence of angle ° on the results.
ocities (if measured in the upper 25 metres) would imply Figure 9.12 from Capozza, 1977, shows not only the
low Lugeon values (0.33 to 0.01 Lugeon) if one accepts higher velocity after grouting but also the increased
L  1/Q as a useful lower bound estimate, as discussed anisotropy, which was closer to that of the unweathered
172 Rock quality, seismic velocity, attenuation and anisotropy

formation at depth. Presumably, the weathered and


jointed sandstone layers between the marl were more
easily injected, giving this increased anisotropy.
Wenhua, 1991, described the use of seismic velocity
measurements to monitor the effects of grouting on the
deformation modulus of jointed and faulted power-
house foundations, at the 1750 MW Gezhouba hydro-
electric project on the Yangtze River in China. The
fracture zone of concern originally had a modulus of
deformation as low as 0.1 GPa and a permeability
greater than 10 Lugeons, and was affected by small,
karstic voids and cracks. Velocity measurements in
the faulted and permeable zone, showed values of
2.5–3.2 km/s after grouting, compared to 1.5–2.0 km/s
before the treatment. The average values for the whole
foundation were 3.18 km/s before grouting and
4.74 km/s after grouting which imply an effective
Q-value increase from (very approximately) 0.5 to 17 or a
Lugeon value reduction from perhaps 2 to 0.06 (using
equations 9.1 and 9.2). This implied low Lugeon value
is so low that a stress related effect on the 4.74 km/s
velocity achieved after grouting is suspected.
As suggested earlier, the stress related effect may be a
combination of depth (greater than the reference 25 m,
Figure 9.4) and post-stressing by high pressure grout-
ing, which could give a ‘locked-in’ stress after curing at Figure 9.13 Top left: a depiction of a Lugeon or water injection
least in vertical or sub-vertical planes. In horizontal test in a rock mass with three joint sets, and the Snow,
planes any potential ‘locked-in’ stress caused by local 1968 idealized cubic network, consisting of a Poisson-
‘lenses’ of grouting, would need to be over a limited distributed, and limited number, of conducting
volume, if arching were to give a local stress greater ‘smooth parallel plates’ with equal permeability. The
than v. ‘Artificially’ high Vp values could then be regis- lower diagrams emphasise the joint-roughness-related,
tered, which might exaggerate the true effect of the inequality of the physical joint apertures (E), and the
grouting, which is primarily designed to reduce permea- theoretical hydraulic apertures (e). Barton, 2004a.
bility, but has several other positive effects (i.e., increased
modulus, shear strength, etc.).
beneath environmentally sensitive areas, where ground-
water draw-down cannot be tolerated.
9.6 Interpreting grouting effects in Most tunnel engineers experience that correctly
relation to improved rock mass carried out pre-grouting reduces leakage, and that it
Q-parameters apparently increases deformation modulus and prob-
ably shear strength, since tunnels that are pre-injected
Since we have indicated a general potential relationship show each of these implied characteristics, meaning
between Lugeon value, Q-value, and measured velocity, improved stability, less deformation, and lessened sup-
and have also obviously noted an apparent change in port needs. The same is probably true in dam founda-
Lugeon value and velocity as a result of grouting, it is of tions, minus the support needs.
interest to investigate the potential physical effects of A helpful, if very idealized figure, concerning the
grouting on Q-parameters. The tunnelling situation ‘available’ joint porosity for potential grout penetration,
can be used in this exercise, as pre-injection is a com- is given in Figure 9.13. The top right-hand diagram is
monly needed measure to reduce problems ahead of a based on Snow, 1968 with the addition of non-
tunnel face, either in permeable, leaking rock masses, or conducting joints between Snow’s idealized cubic
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 173

Since 1 Lugeon  107 m/s, and 107 m/s  1014 m2,


a laminar flow 3D interpretation of Lugeon tests can be
expressed as follows:

e  (6LS  108 )
1/ 3
(9.8)

where (e) and (S) in millimeters, L is the average


Lugeon value, and each apply to the local domain, rock
type, or borehole depth.
Although average physical apertures (E) are signifi-
cantly larger than (e), they are hardly of different orders
of magnitude. It is therefore clear that there will be dif-
ficulties of grouting a Vp  4.5 km/s rock mass (meas-
ured at nominal 25 m, shallow refraction seismic depth),
if we assume the approximate validity of equation 9.2:
L  10(3.5V ), suggesting a Lugeon value of only about
p

0.1 in this case.


It is simple to understand from grouting case records
Figure 9.14 The evaluation of equation 9.8 in graphic and tabu- that the higher the Lugeon value, or the lower the vel-
lated format, for typical Lugeon values between 0.01 ocity before grouting, the better the potential improve-
and 100, equivalent to assumed isotropic rock mass ment (e.g. Figure 9.11). A useful demonstration of this
permeabilities of approx. 109 to 105 m/s, and aver-
is the practical case of trial grouting of a dam abutment
age spacings for the water-conducting joints of 0.5 m
in Figure 9.15, from Quadros and Correa Filho, 1995.
to 3.0 m.
Three boreholes permeability-tested before grouting,
were re-drilled in roughly the same location following
network. The physical (aperture E), compared to the the grouting, so that the before-and-after permeabilities
theoretical (apertures e), available for grouting are could be compared. The lower-left diagram, treating
depicted in the lower diagrams. just the individual borehole results, suggested that the
A further visualization of the size of the theoretical following before-and-after results could be expected.
hydraulic apertures available for grouting, if the rock (Only industrial cement was used in these tests).
mass had three equal joint sets, is given in Figure 9.14,
which was derived in Barton, 2004a, from equations in Before After
American units from Snow, 1968. Based also on the k  103 m/s k  107–108 m/s
hydraulic theory of Louis 1967: k  105 m/s k  106–107 m/s

In the tunnel situation, the need for reduced tunnel


1. Permeability of one smooth parallel plate:
support following pre-grouting, can be documented, if
prognoses of required support using an ‘ungrouted Q’
e2
k (9.5) are accepted as realistic. This claim has been supported
12 by recent rail tunnels for the Norwegian Jernbaneverket
in the Oslo area. Tunnels were driven under built-up
2. Permeability of 1 set of parallel plates: areas founded on clays, using over-lapping pre-
injection ‘umbrellas’ established every 3 to 4 rounds, by
e2 e performing a regular, high pressure (5 to 10 MPa),
K1   (9.6)
12 S single-stage, 24 hours-duration pre-grouting routine,
over many kilometres if tunnel. (Moen, 2004).
3. Permeability of ‘the conducting rock mass’ ( 3 sets): Since tunnel deformation is closely linked to SPAN/Q
(Barton et al., 1994, Barton, 2002) and support needs
2e 3 are linked directly to Q, the inescapable conclusion
! mass  (9.7)
12S (which would also be arrived at by velocity monitoring
174 Rock quality, seismic velocity, attenuation and anisotropy

Figure 9.15 Permeability testing at a dam abutment in Brazil, using before-and-after testing of the effect of grouting, and both single-hole
and 3D measurements. Note the rotation of permeability tensors, and their reduced magnitude, suggesting progressive sealing
of joint sets. This suggests a possible scenario for individual Q-parameter improvements. Quadros and Correa Filho, 1995.
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 175

and deformability testing) is that the effective Q-value 30 1.5 0.5


Q1     0.8 (9.10)
has itself been increased by the pre-injection. 15 2 1
The Q-value (Barton et al., 1974, Barton and
Grimstad, 1994) is determined from the modified core
60 24 0.66
recovery RQD (i.e., counting competent pieces of Q2     9  23 (9.11)
core  100 mm in length as recovery). In addition to 9 1  0.75 1
RQD, Q is calculated from the number of joints sets
(Jn), the roughness (Jr) and the degree of alteration (Ja) The effective Q-value has increased in terms of round
of the least favourable set, and from the water inflow figures, by a factor of 10 to 30, which is broadly consist-
(Jw) and stress/strength condition (SRF). From Figure ent with the increased Vp and M values, and with the
9.4, a velocity increase of 1 km/s from say 3.5 to reduced Lugeon value and rock support needs. It
4.5 km/s at a dam site, or in a wet, jointed zone ahead should be noted in particular, that when the seldom
of a large tunnel, will imply that the Q-value has reported or measured 3D permeability is analysed in
increased from 1 to 10 as a result of grouting. before-and-after-grouting scenarios, a rotation of the
Following equation 9.1, a drop in Lugeon value from permeability tensors (and reduction of their magni-
1.0 to 0.1 is also implied, and using equation 9.3: tude) is seen (Figure 9.15, from Quadros and Correa
M  10Q1/3 c , the modulus of deformation may be pre- Filho, 1995). This is the tentative justification for sug-
dicted to have increased from 10 GPa to at least gesting, as above, that the least favourable joints – and
20 GPa. Are these changes possible to explain via those causing the lower before-grouting velocities – are
changes in the six component Q-parameters? The those that are (first) sealed by the grout.
answer is definitely yes, but the exact answer will always In this particular example we can estimate the fol-
be unknown. lowing ‘hard rock, shallow near-surface’ results for
We can speculate that the following orders of magni- before and after grouting, based on Vp  3.5  log10
tude of effective rock mass quality improvement (each Q, L  1/Q, M  10Q1/3:
very modest), may occur in practice during grouting
ahead of a tunnel or at a dam foundation:
Table 9.2 Potential effects of grouting according to empirical
1. RQD of say 30%, increases to say 60%, due to predictions.
grouting of the most prominent set of joints that
Before grouting After grouting
were most permeable. This occurs largely as a result
of hydraulic joint jacking. Q 0.8 9 → 23
2. Jn of say 15 (four sets) is effectively reduced to 9 Vp 3.4 (km/s) 4.5 → 4.9 (km/s)
(three sets) for the same potential reasons as above. L 1.3 (Lugeon) 0.1 → 0.04 (Lugeon)
(This is a very conservative argument). M 9.3 (GPa) 21 → 28 (GPa)
3. Jr of 1.5 (rough, planar) changes to 2 (another set)
or to 4 (discontinuous), also for the same reasons.
4. Ja of 2 (weathered) changes to 1 (another set) or As with some of the cases reviewed earlier, this appar-
to 0.75 (cemented), also for the same reasons. ently good grouting result would need to be attributed
5. Jw of 0.5 (high pressure inflow) changes to 0.66 to hydraulic joint jacking and perhaps to the use of
(small inflow) due to preferential sealing of the micro-cements. In relation to the Inguri arch dam
most permeable set. (This is also conservative). (Figure 9.11), Savich et al., 1983, would allocate the
6. SRF of 1 (unchanged). (In the case of a minor fault, result (Vp  3.4 → 4.5 → 4.9 km/s) to class II (good
even SRF might change). grouting result).
The interaction of rock mechanics, rock hydraulics and
We therefore have the following potential ‘before’ and rock dynamics through application of seismic monitor-
‘after’ scenarios: ing and rock quality description has many applications
From: for rock engineers. The ‘core’ interactions (Vp, Q, L
and M) illustrated above and in Figure 9.4 can also be
RQD J J
Q   r  w (9.9) expressed in alternative ways. By ‘extracting’ the uni-
Jn Ja SRF axial strength ( c) of a rock from Q c( Q  c/100) we
176 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(b)

(c)

Figure 9.16 Alternative nomograms for estimating typical interactions between c, Q, M, Vp and L. Note that porosity and depth (or stress)
effects have been ignored for simplicity of presentation. (All predictions for nominal n  1% porosity, and typical refraction
seismic depth  25 m.)
Relationships between Vp, Lugeon value, permeability and grouting in jointed rock 177

can make c one of the principal variables as shown in GPa and Vp are km/s respectively. The common x-axis
the nomograms for M, L and Vp shown in Figure 9.16. in each diagram is the Q-value of the rock mass, i.e., the
The numbers distributed within Figure 9.16a are rock mass quality and not the seismic quality, although
estimates of the static deformation modulus (M), while as discussed in more detail in the next chapter, these
the numbers distributed within Figure 9.16b and c, are numbers, each having the same Q symbol, are
the estimates for P-wave velocity. The units of M are inevitably related quite closely.
II Introduction to Part II

The subduction zone diagram showing ‘extremely low called ‘Q’ (range 0.001 to 1000) and seismic quality Q
Q’, ‘low Q’ and ‘high Q’ reproduced on the back cover (range 1 to 5,000), where attenuation  Q1, could in
of this book, the ‘familiar’ Vp-depth trends of mid- some way be related, proved to be one of the incentives
ocean ridge seismic investigations, and the continent- for deeper research into seismic phenomena, and are
wide seismic velocities also showing ‘familiar’ increase the reasons for developing Part II material.
with depth were each strong reasons for delving deeper Part II contains a wide sampling of interesting large
into the subject of ‘seismic velocity and rock quality’, scale continental and sub-ocean seismic behaviour, also
which was the original title planned for this book. high pressure laboratory rock physics tests designed to
Possible parallels with engineering scale phenomena improve understanding of both crustal and reservoir
were evident, in something resembling a ‘fractal’ earth. variation-with-depth phenomena. A broad sampling of
Of course the subduction zone ‘Q’ values proved to in situ reservoir related topics is given in later chapters,
be attenuation related. However the possibility that a such as borehole stability and their seismic effects, and
commonly used rock mass quality parameter, also fractured reservoir investigations, involving P-wave
anisotropy, S-wave splitting and polarization, and poro-
elastic modelling of the dispersive and anisotropic
nature of fractured reservoir simulations. An attempt
has been made to bridge between engineering, geologi-
cal and geophysical scales of depth and time, in this con-
tinued investigation of ‘seismic velocity and rock quality’.
The last chapter addresses geomechanics understand-
ing of joint and fracture behaviour, in particular perme-
ability-stress performance, with a view to suggesting
alternative interpretations of aligned fracture orienta-
tions that actually involves multiple sets. Deep well
behaviour in which impermeable and permeable fracture
sets are separated by the determination of either domi-
nance of normal stress or dominance of shear stress, rep-
resents a more correct understanding for maintenance of
Fig. PART II Schematic section, after Barazangi and Isacks 1971 permeability in the face of high effective reservoir
and Kearey and Vine 1996, of the Tonga arc, with stresses. The ‘parallel to H max’ assumption for aligned
inferred seismic Q variations. A possible relationship single sets of conducting fractures from shear-wave
between seismic Q and rock mass quality Q is one of anisotropy may be an over-simplification, and is often in
the first objectives of Part II. conflict with geomechanics test data and theory.
10 Seismic quality Q and
attenuation at many scales

In this chapter the term ‘Q’ used in the title in the clas- in the same volume. Common sense would suggest that
sic paper of Knopoff, 1964 will be distinguished from Qseis can never be less than 2; however values below
the engineering rock mass quality Q-value of Barton this magnitude are quite frequently recorded near the
et al., 1974, by reference to the seismic quality as ‘seis- surface, including negative values which presumably
mic Q’, Qseis, or Qp or Qs if the compressional wave or may reflect interpretation difficulties of some sort.
shear wave components have been distinguished. In At the time of Knopoff’s review it was customary to
fact, as we shall see, there are obvious connections assume that Qseis was substantially independent of fre-
between Qseis and the rock quality Q-value; a heavily quency. His assumptions of ‘a homogeneous sample’ and
jointed clay-bearing rock mass with low Q-value (prob- ‘at low frequencies’ are clearly important in view of what
ably less than 0.1) will inevitably cause great attenuation is now understood about potential dissipation mechan-
and have a correspondingly low Qseis (perhaps less than isms in microcracked rock samples or in rock masses with
5), while an almost unjointed massive rock mass with sets of bedding planes and/or joints. Laboratory experi-
very high Q-value (e.g. 100–500) will inevitably cause ments on many homogeneous solids had shown that up
little attenuation and have correspondingly high Qseis, to moderately high frequencies, the dimensionless quan-
depending on whether shallow or at great depth. tity Qseis was virtually independent of frequency. This
Knopoff, 1964, introduced his review of seismic Q (or preliminary conclusion indicated that the mechanism by
Qseis) by stating ‘Were it not for the intrinsic attenuation which energy was removed from elastic waves in solids
of sound in the earth’s interior, the energy of earthquakes was not the same as the mechanism for attenuation in liq-
of the past would still reverberate through the interior of uids, where attenuation is frequency dependent.
the earth today. The chaos resulting from this awesome Some typical values of Qseis for longitudinal excita-
prospect is a speculation which lies outside the scope of tion of various solids, selected from Knopoff, 1964, are
this paper.’ We can conclude that Qseis and any of the reproduced below.
physical reasons for Qseis that are captured in the Q-value In this very selective list, the attempt is made to link
rating (Appendix A) are fundamental to our well-being, Qseis to the relative stiffnesses of these materials. In reality
even though low values of both may cause problems when the satisfactory-looking ‘order’ seen here is more scat-
tunnelling or when preparing a large dam foundation. tered. One may comment already that the sandstone,

Table 10.1 Some examples of Qseis for longitudinal or bending


10.1 Some basic aspects concerning excitation of various solids, selected from Knopoff,
attenuation and Qseismic 1964, sorted by magnitude.

Material Qseis
Using the definition of Qseis given by Knopoff, 1964, as
a starting point, we may refer to the familiar electrical Steel 5000
circuit theory for energy loss: Copper 2140
Silica 1250
Glass 490
2 E Diorite 125
 (10.1)
Q seis E Limestone 110
Lead 36
Sandstone 21
In this definition, E is the amount of energy dissi-
Shale 10
pated per cycle of a harmonic excitation in a certain
Celluloid 7
volume, and E is the peak elastic energy in the system
182 Rock quality, seismic velocity, attenuation and anisotropy

since presumably not jointed, was likely to have been frequency dependent, due to the inertial forces of the
very weak and porous, in view of later values for sand- fluid in the microcracks and joints, and due to scattering.
stones that we will review. Walsh, 1966, proposed a frictional-dissipation-at-
Interestingly, and as a less serious aside, the 2003 crack-surfaces model to explain the simpler attenuation
Paramont film ‘The Core’, about an improbable voyage in dry rock. There were parallels with his observation of
towards the centre of the earth to ‘fix’ an electro- hysteresis when loading and unloading rock in uniaxial
magnetic hazard, contains an opening sequence where compression.
the soon-to-be-seconded professor (Dr. Joshua Keys, Concerning attenuation in intact rock, Walsh envis-
played by Aaron Eckhart) has written the following aged the following. Among the large number of cracks
Q-quality factors on the blackboard of the University of all orientations and lengths, some are open and some
of Chicago lecture room: are closed at any given pressure. As a compressional
wave traverses the rock, (micro-scale) sliding on one
Shale 20 to 70 crack-face past the other will occur on cracks which are
Limestone 45 to 90 barely closed and which have favourable orientation
Granite 40 to 230 with respect to the wave propagation. This crack-face
Also: Q : ": no attenuation motion is opposed by friction, and some of the elastic
energy of the wave is dissipated.
The writer was informed about this sequence by a As the wave traverses the material, the normal stress
lawyer who wondered if it was ‘my’ Q. These Qseis mag- between the crack-faces increases, and thus the fric-
nitudes are of course entirely feasible extensions of the tional shear stress also increases. As the wave passes, the
above list of Qseis magnitudes selected from Knopoff. direction of the frictional shear stress is reversed, and
As we will see, the numbers for any rock will change again work must be done against friction as the crack
with degree of microcracking, with pressure, with the returns to its equilibrium position. Clearly, micro-scale
dry or saturated state, with weathering, and with the deformations are implied here.
degree of jointing and faulting when at larger scale. Numerous mechanisms have been proposed to
Due to each of the above, the frequency band of the explain attenuation of seismic waves in rock and in rock
dynamic loading will also affect the result, due to fac- masses. Johnston et al., 1979 listed the following in
tors to be explored in this chapter. their landmark paper:
It is now known that higher frequencies are attenuated
at a higher rate than lower frequencies. Thus in a con- ● Matrix anelasticity
stant seismic Q (or rock quality Q) region of the rock ● Frictional dissipation due to relative motions at
mass, the amplitude of high frequency waves will grain boundaries and across crack surfaces (cf.
decrease faster than that of the low frequency waves. Walsh, 1966)
Although most of the early evidence suggested that the ● Fluid flow causing relaxation due to shear motions
seismic quality factor Q was frequency-independent over at pore-fluid boundaries
a wide range of frequencies (e.g. 102 to 107 Hz for the ● Relative motion of the matrix frame with respect to
case of shales, MacDonal et al., 1958), this is now gener- the fluid inclusions in the case of fully saturated
ally rejected, as a result of laboratory tests conducted rock (cf. Biot, 1956a)
under different frequencies, under smaller more realistic ● Squirt phenomena (cf. Mavko and Nur, 1975 and
strain levels and over wider ranges of confining pressures. O’Connel and Budianski, 1977)
More recent data from in situ well tests conducted ● Gas pockets squeezing when only partial saturation
over wide ranges of frequency show the fundamental ● Geometrical effects due to small pores, larger irregu-
frequency dependence of seismic Q, due to the range of larities, thin beds (this category obviously extends
scales of the various attenuation mechanisms. Some of to major discontinuities, faults, rock boundaries,
these newer sets of data will be reviewed in this chapter. dykes etc.)
Although Knopoff, 1964 assumed that the attenu-
ation of elastic waves in dry (intact) rock was inde- Attempts to illustrate some of the smaller scale
pendent of frequency, it is a different matter when mechanisms of intrinsic attenuation are reproduced
microcracks and joints and water saturation (or partial from Johnston et al., 1979, in Figure 10.1. The fluid
saturation) are added. Attenuation increases and becomes flow attenuation mechanisms really fall into two
Seismic quality Q and attenuation at many scales 183

Figure 10.2 A generalized chart concerning strain magnitudes and


frequencies, for various deformation processes. After
Batzle et al., 2005.

Figure 10.1 Schematic illustrations of several of the proposed intrin-


sic attenuation mechanisms. Johnston et al., 1979.

frequency-dependent categories: the inertial resistance


which is important at ultrasonic frequencies, and squirt
flow which is more prominent at lower frequencies.
It appears that friction across thin cracks and grain
boundaries may be the dominant attenuation mech- Figure 10.3 Conceptual diagram of elastic constants for different
anism (at small scale), if strain levels are sufficient. relaxation mechanisms, with a frequency scale, and an
Increasing pressure decreases the aperture and effective indication of whether the fluid involved is of low or
high mobility. After Batzle et al., 2005.
number of cracks, and thereby reduces attenuation.
Water wetting and saturation reduces the friction coef-
ficient, thereby increasing attenuation. We will see Figures 10.2 and 10.3. These supplemented their
much more detail of these aspects in the next section, extended abstract. They give a useful perspective on the
and a detailed treatment of the effect of strain levels orders of magnitudes involved in these fundamental
and pressure on attenuation mechanisms. earth-science topics. The authors emphasised that
moduli or velocities measured in one amplitude
or frequency domain were usually not valid in other
10.1.1 A preliminary discussion of domains, since different deformation mechanisms
the importance of strain would likely be operating.
levels Concerning the ‘elastic 4 plastic’ ‘static 4 dynamic’
cross-plot in Figure 10.2, one may put forward a
A useful summary of key concepts concerning elastic ‘jointed-rock-mechanics’ viewpoint that in the presence
(contra plastic) strain as a function of frequency, and of the usual heterogeneities of jointed, stressed rock
relaxation mechanisms as a function of frequency, was masses, there will be a tendency for the four ellipses
given by Batzle et al., 2005. Two diagrams presented in below the horizontal ‘static-dynamic’ axis to stretch
their EAGE Madrid poster are reproduced here, in their long axes upwards, into larger strain territory. The
184 Rock quality, seismic velocity, attenuation and anisotropy

reason for this opinion is that dynamic joint compli- point for the similarly steep, or much increased gradients
ances derived from the registration of seismic of 1/Z. It is surely logical to assume that the four ellipses
anisotropy, reviewed in later chapters, have inverted depicted in Figure 10.2 stretch more into higher strain
magnitudes that are, perhaps surprisingly, partly within territory, the lower the rock quality, with the likelihood
experimental ranges of pseudo-static joint stiffness data, of some slight, irreversible deformation in the case of seis-
from rock mechanics ‘macro-deformation’ testing. mic, low frequency motion, especially near the rock sur-
It is difficult to believe, on this basis, that joint micro- face, where stresses and deformation resistance are low.
displacements involved in developing the characteristic According to the dispersion relation of Kjartansson,
deformation-load units of compliance (m.Pa1) from 1977, cited by Mavko and Nur, 1979, the attenuation
in situ seismic inversion, could be as small as ‘sub- Q1, or inverse seismic quality Q can actually be used
atomic’, as a prominent physicist has suggested. Such an to explain the difference between the static and
opinion probably stems from consideration of the effect dynamic moduli of intact rock (See Chapter 6 for gen-
of microstrain on microcracks (in intact samples), giving eral results for rock masses). It is well known that the
‘too small’ influence to mobilize conventional concepts dynamic modulus can be at least double that of the
of friction, as we shall see shortly. static modulus even in intact rock, if flat pores or
In the in situ reality, a rock mass consists of joints and microcracks are present. Part of the difference in mod-
discontinuities with both length dimensions, and spa- uli may be due to fluid stiffening, in addition to the
cings, many, many orders of magnitude larger than the above dynamic-compliance/static-stiffness differences
rock physicists microcracked ‘intact’ samples. The nano- required when going up to in situ scale.
strains to sub-microstrains presumably experienced in a According to Kjartansson, 1977, the ratio of moduli
rock mass during the passage of seismic waves (depend- (M) at different frequencies (f ) can be expressed as:
ing on distance from source and its magnitude) may 2
then, through discrete micro-displacement in the rock M1 f  pQ

  1  (10.2)


mass, experience attenuation due to the larger scale fea- M2  f 2 
tures as well.
Possibly the response of the joint to the dynamic pulse, It is assumed here that Qseis remains roughly constant
is to initiate response from the current operating normal over the frequency band of interest. Taking a ‘static’
and shear stress-deformation gradients. Dynamic micro- value of f2  0.01 Hz and a dynamic f1  105 Hz
excursions above and below these gradients would then and Qseis  1000, 100, 50 and 20, we find predicted
occur, with an increased (or nearly equal) slope presum- ratios of M1/M2 of 1.01, 1.1, 1.2 and 1.7 respectively.
ably depending on the quality of the joint walls and sur- Obviously, the more flaws (pores, microcracks) that are
rounding rock. present in a rock sample, the lower will be the seismic
A useful geophysics concept, in this context, is that quality Q, and the higher the predicted ratio of M1/M2
‘rock quality is defined as whether Edynamic is more than or (the dynamic/static ratio of moduli).
equal to Estatic’. Clearly, as shown in Chapter 6, when rock The above is consistent with the idea of a broadly
is hard and joints are fresh, the inequality of these two related Qseis and rock quality Q, since dynamic moduli
moduli is small – in which case the inverse of ‘dynamic’ diverge more from the static moduli, as rock (mass)
joint or fracture compliance is also likely to be similar to quality reduces. The above ‘intact rock’ difference is
the ‘static’ joint or fracture stiffness. When on the other accentuated when larger scale is considered, since the
hand rock (and joint) quality is poor, which in the rock rock joints will usually have lower values of the ‘static’
mechanics world would be when rock quality Qc was low normal and shear stiffnesses, than the inverse of the
( and joint wall compression strength JCS also low, per- dynamic compliances of the same joints.
haps even with clay-smear-or-filling), there will then be a
big inequality, with Edynamic  or  Estatic, and a pre-
sumably corresponding inequality of 1/dynamic compli- 10.1.2 A preliminary look at the
ance  or  ‘static’ stiffness, or as we shall see in attenuating effect of cracks
Chapter 16, 1/ZN # Kn, (and 1/ZT  or  Ks). of larger scale
In both the above scenarios: high or low rock quality,
and the existing stress-deformation gradients, be they A useful insight into the effect of changes in crack
steep or shallow, will likely determine the ‘static’ starting porosity (and number of cracks) on the seismic quality
Seismic quality Q and attenuation at many scales 185

factor Qseis, was given by Remy et al., 1994. We will bedding planes (during the 8th cycle). Physical evidence
utilise this in this introductory section, before review- for the cracking was seen from hydrostatic loading tests
ing intact laboratory data concerning seismic Q. The on the cubic samples, where definition of the total volu-
authors’ laboratory investigations involved sixteen metric crack porosity (the sum of the components of each
freeze-thaw cycles (20°C to 20°C) over a period of axis) was recorded. This parameter increased successively,
sixteen days, in order to simulate part of the first appear- and uniformly, during sixteen cycles of freezing and thaw-
ance of weathering effects. The rock investigated was a ing, and clearly intimately affected the reduction in Qseis.
thin-bedded (1 cm), Jurassic limestone from Lorraine The reduced values of Vp and Qseis with successive
in France. Cylinders (5 cm diameter, 10 cm length) and accumulation of crack-related damage have direct par-
cubes (5 cm sides) were used, having a bulk density of allels in rock mass quality changes (i.e. reduced rock
2.1 gm/cm3 and porosity of 22%. The bedding planes mass Q-value due to the fact that RQD reduces, Jn may
were perpendicular to the axes of the cylinders, and par- increase, Jw reduces and, subsequently Ja increases as a
allel to the top surface of the cubes. The repeated cycles result of weathering. See Appendix A for descriptions of
of freezing (5 hrs), frozen (6 hrs), thawing (5 hrs), the Q-parameters of rock quality (Barton et al., 1974).
thawed (8 hrs) and corresponding changes of P-wave
velocity are shown in Figure 10.4.
P-wave velocities were higher when frozen (e.g.
4.7 km/s) than when thawed (e.g. 3.4 km/s) due to the
higher wave velocity in ice (3.8 km/s). Maximum veloc-
ities were reached at the end of the freezing. As shown
in Figure 10.5a and b, Vp (frozen state) fell with each
cycle, while Vp (thawed) fell most rapidly on the first
two cycles. It is important to note that the creation of
new cracks caused under-saturation of the initi-
ally water-saturated samples, which were jacketed, and
immersed in a solution of methanol. Figure 10.4 A unit freeze-thaw-time cycle of 24 hours applied to
The two marked drops in Qseis values signify cracking thin-bedded limestones, and its basic effect on Vp.
episodes, the second of which was perpendicular to the Remy et al., 1994.

Figure 10.5 a) Velocity Vp versus number of freezing and thawing cycles. b) Seismic Q versus number of freezing and thawing cycles. (Remy
et al., 1994).
186 Rock quality, seismic velocity, attenuation and anisotropy

In the literature there are numerous references to the when considering both the P-wave and S-wave related
relatively low values of Qp (and Qs), in near-surface attenuations, the ratio of Qs/Qp proves to be an even
jointed or altered rock (e.g. 50 or less) and the higher better indicator of the degree of saturation.
values for intact samples of rock (e.g. 10 to 250) and In each case, pressure, with its microcrack-closing abil-
the higher still values for deep igneous and metamor- ity, causes a rise in velocity and a reduction in attenuation.
phic basement rocks (e.g. 100 to several 1000s), sills Qseis therefore rises. According to the model of attenu-
(e.g. 2000) and salt (e.g. 1000). ation developed by Johnston et al., 1979, the relative
contributions of friction and fluid flow on the overall
attenuation are as shown in Figure 10.8.
10.2 Attenuation and seismic Q from
laboratory measurement

A compilation of Qseis values for specific groups of rocks,


without distinction between different frequencies or
degrees of saturation shows, inevitably, a wide scatter as
shown in Figure 10.6, from Bradley and Fort 1966. This
shows only porosity as the plotted variable, and conse-
quently a range of Qseis from less than 10 (for porous
sandstones) to nearly 900 (for low porosity igneous and
metamorphic rocks). There is a general trend of Qseis
inversely proportional to porosity, but this is comprom-
ised by too many hidden mechanisms of attenuation.
Early investigations of the effect of the degree of
water saturation in reducing Qseis for porous rocks, and
Figure 10.7 Saturation and pressure dependence of Qseis. From
its strong pressure sensitivity as the rock reverts from Gardner et al., 1964 data.
dry to different degrees of saturation, are shown in
Figure 10.7, from Gardner et al., 1964 data, repro-
duced by Johnston et al., 1979. As we shall see later,

Figure 10.6 Qseis as a function of porosity for igneous and meta-


morphic rocks (triangles), limestones (squares), and
sandstones (circles). A wide range of frequencies and
degrees of saturation contribute to the scatter of data. Figure 10.8 Relative contribution of friction-based and fluid flow-
From Bradley and Fort 1966, reproduced by Johnston based attenuation for a brine saturated Berea sand-
et al., 1979. stone, according to the model of Johnston et al., 1979.
Seismic quality Q and attenuation at many scales 187

As pointed out by Johnston et al., 1979, since the The question of whether friction is a viable source of
porosity and permeability (of these intact specimens) is seismic attenuation; along microcracks, across crack-
relatively unchanged by the range of pressures applied, tips, (and also along joints and filled discontinuities,
there is limited effect on the fluid-flow contribution to and within the multiple surfaces of faults), will now be
attenuation. Such would presumably not be the case if addressed again, with the benefit of more understand-
a jointed specimen or a jointed rock mass was involved, ing of the effects of strain levels, provided by Winkler
where pressure sensitivity of the permeability and sec- and Nur, 1982. With its title: ‘Seismic attenuation:
ondary porosity would be marked, and non-linear, effects of pore fluids and frictional sliding’, one would
thereby giving a strong rise in Qseis with the reduced certainly expect that both mechanisms were still to be
attenuation and velocity increase. emphasised as potential sources of attenuation.
In their conclusions the authors however, state the
following: ‘Since the conditions required for sliding
10.2.1 A more detailed discussion of friction to be observed (large strains and small confin-
friction as an attenuation ing pressures) generally do not apply to seismic wave
mechanism propagation in the earth, we conclude that simple fric-
tional sliding is not a significant attenuation mech-
According to the models of Johnston et al., 1979, the anism in situ.’
relative effects of frequency and pressure can be com- Their conclusion was drawn, at least partly, on the
bined to elevate the total Qseis (specifically Qp) of the basis of extensional resonance tests, conducted on long,
Berea sandstone. At low pressures, the friction mech- thin (intact) bars of homogeneous rock, such as sand-
anism dominates and is almost independent of fre- stone, which were contained inside a long pressure vessel,
quency. With increasing pressure and low frequencies and made to oscillate with an electro-magnet, while sup-
Qp climbs beyond 100, but as frequency increases there ported rigidly at their mid-point. Figure 10.10 shows the
is a reduction of Qp due to the contribution of squirt results of resonance decay measurements, giving both
flow and so-called shear relaxation. Eventually, at very
high frequencies, Qp declines sharply again due to scat-
tering. This general scheme of predicted behaviour is
illustrated in Figure 10.9.

Figure 10.10 Variation of attenuation (1000/QE) and velocity


with strain amplitude, based on extensional reson-
ance decay measurements on long (intact) bars of
sandstone, suspended at their mid-point in a pres-
sure vessel, and excited by an electro-magnet at one
end, with a phonograph pick-up at the other end.
(Note Q⫺1 sensitivity of 19%, and velocity sensitiv-
Figure 10.9 Total Qp predicted for brine-saturated Berea sand- ity of only 0.7% to the 2-order of magnitude strain
stone, from Johnston et al., 1979. amplitude variation). Winkler and Nur, 1982.
188 Rock quality, seismic velocity, attenuation and anisotropy

Table 10.2 Effect of strain amplitudes on extensional seismic


quality, showing negligible effect on velocity. Selected
data from Winkler and Nur, 1982.

Material Strain amplitude QE Velocity

Sierra white 1.44  106 185 3,629 m/s


granite 4.15  108 204 3,637 m/s
Berea 2.10  106 103 1,937 m/s
sandstone 2.30  108 140 1,955 m/s
Lucite 1.43  106 23.4 2,108 m/s
8
3.04  10 23.2 2,108 m/s

velocity and Q1 as a function of strain amplitude. The


authors used frequencies from 500 to 9000 Hz, and
studied the effects of confining pressure, degree of sat- Figure 10.11 Effect of confining pressure in reducing the strain
uration, strain amplitude, and frequency. amplitude-dependence of extensional attenuation
The changes in attenuation and velocity they observed QE, for dry Berea sandstone. Curve A  1 MPa, B 
with increasing strain amplitude (Figure 10.10) were 2 MPa, C  3 MPa, D  5 MPa, E  5 – 3  2 MPa
interpreted as evidence of frictional sliding at grain (helium pore pressure of 3 MPa). Note therefore the
contacts. But since this amplitude dependence suppos- closeness of curves B and E. Winkler and Nur, 1982.
edly disappeared at strains and confining pressures that
they considered were typical of seismic wave propaga- 106, as shown when comparing Figures 10.10 and
tion in the earth, they consequently inferred that fric- 10.11, the authors suggested that it was significant that
tional sliding was not a significant source of seismic this strain level was in the same range as that needed for
attenuation in situ. cusped stress-strain loops to become elliptical (Brennan
They referred to other problems with the frictional and Stacey, 1977). They then posed the question: why
attenuation mechanism. Savage, 1969 had pointed out is a strain of 106 or larger needed to cause frictional
that for typical strain amplitudes of seismic waves, and attenuation? (At least for the case of intact rock specimens,
for reasonable microcrack dimensions, the computed excited in the extensional mode?).
slip across crack faces would be less than the inter- They explored the answer to this question by sug-
atomic spacing. They assumed, probably correctly, that gesting that displacements across crack surfaces should
this extremely small interaction would not be described at least be comparable to inter-atomic spacings of about
by conventional models of macroscopic friction. In 1010 m. They then equated a shear strain (␧) to a max-
addition, they referred to the widely held assumption imum displacement (d ) across a crack of length (L),
that frictional attenuation caused nonlinear wave propa- suggesting d  ␧.L. With (d )  101 0 m, (␧)  106,
gation, which had not apparently been observed at the a crack length (L) of 104 , or 0.1 mm is implied. The
low strain amplitudes typical of seismic waves. authors considered this to be a realistic upper-bound for
Winkler and Nur provided a useful summary of the microcrack sizes in rock, so concluded that at strains below
strain amplitude dependence of extensional attenuation 106, sliding displacements would generally be too small
QE for intact samples of several rock types, and some for friction to describe the (sub-micron) interaction.
man-made materials. (The Massilon sandstone result is The obvious corollary to this is to pose the question:
shown in Figure 10.10). Only the materials (i.e. rocks) what about all the larger cracks, i.e. intra-bedding joints,
that contained potential (micro) sliding surfaces, indi- tectonic joint sets, major clay-filled discontinuities,
cated strain amplitude dependence, and the author’s and multiple internal interfaces in fault zones, all of
tests showed that (intact) rock samples almost lost this which have large, or extremely large (L). With potential
dependence with moderate increases of pressure, as length dimensions of 0.1 m 1.0 m and 10 m for the
shown in Figure 10.11. ‘smallest’ three of the above five categories of discontinu-
Noting the effect of confining pressure on extending ity, and assumed spacings of the same order of magni-
the strain amplitude ‘limit’ for QE sensitivity, to about tude (for convenience of estimation), an unchanged
Seismic quality Q and attenuation at many scales 189

Vp (km/s)
3

2
Dry
Partially (~90%) saturated
Fully saturated
1

1.5 2.0 2.5


(a) Vp / Vs

Dry
2.0
Figure 10.12 Seismic Q as a function of angular frequency and Partially (~90%) saturated
Fully saturated
shear strain level, measured on cylinders of near-surface
clay, in a resonant column apparatus. Marmureanu 1.5
et al., 2000.
Qs /Qp
1.0
continuum-based shear strain of 106 generated close to
a given seismic source, might well imply maximum
0.5
(close to the source) displacement discontinuity events of
the order of 0.1, 1.0 and 10
m for these three joint/dis-
continuity types, if the continuum strain was converted 1.5 2.0 2.5 3.0
to intermittent discontinuous shearing events with the (b) Vp / Vs

same frequency as their length scale. Figure 10.13 Cross-plots of Vp versus Vp/Vs, and Qs/Qp versus
Can such events be the source of dynamic joint com- Vp/Vs, showing the distinctive effects of the dry,
pliances in geophysics, that have recognisable (nearly partly saturated, or fully saturated states, when using
same order) magnitudes and units, as the MPa/mm these parameter ratios. Winkler and Nur, 1982.
pseudo-static stiffnesses of rock joints that are familiar
to rock mechanics engineers? Shear strains decaying to
one or two orders of magnitude less than 106, further 10.2.2 Effects of partial saturation
from seismic sources, are surely still capable of providing on seismic Q
displacement discontinuities of sufficient magnitude for
frictional attenuation to be a valid mechanism in rock Figures 10.13 a and b, show a useful summary of some
masses, as opposed to intact bars of homogeneous rock. of Winkler and Nur, 1982 work on the effects of the
While on the subject of the importance of strain level dry, partly saturated, or fully saturated state on the
and frequency on Qseis, it is of interest to look at soils, P-wave velocity and its variation with Vp /Vs. A ‘com-
nicely illustrated by the results of Marmureanu et al., panion’ set of data for the moisture-detecting ratio
2000, using resonant column equipment. They tested Qs/Qp versus Vp/Vs is also shown.
cylindrical samples from surface soil layers, applying tor- The S-wave attenuation increases with saturation (Qs
sional and longitudinal vibrations, in studies connected reduces), thus making the ratio Qs/Qp a particularly
with seismic risk mitigation. Figure 10.12 shows seismic sensitive indicator of the degree of saturation, since
Q as a function of shear strain level (%) and frequency, P-wave attenuation, though increasing with initial sat-
almost showing independence from frequency over a uration levels, eventually reduces to less than the S-wave
typical engineering seismology range of interest, i.e. attenuation: thus the ratio Qs/Qp reduces to low levels,
about 5 to 100 Hz. The angular and shear strain depend- since Qp has increased. The separation of data into
ence of soil, giving non-linear behaviour, was emphasised ‘environmental compartments’ is very interesting, and
in their focus on earthquake hazard estimation. also useful for in situ interpretation.
190 Rock quality, seismic velocity, attenuation and anisotropy

As we have seen above, seismic attenuation in partly


saturated rock with pore space and microcracks pro-
vides interesting insights into the frequency depend-
ence of seismic Q and Q1 for this small-scale element
of rock mass. The early work of Knopoff, 1964, showed
that attenuation in dry rock was independent of fre-
quency even over a wide range, because of the assumed
velocity-independence of sliding against friction across
crack faces. When a sample is fully saturated, attenu-
ation becomes frequency-dependent because of energy
losses from viscous dissipation, which depends on
shearing velocity.
Figure 10.14 Attenuation as a function of frequency in dry rock
As discussed by Walsh, 1995, when attenuation is (open squares) and rock fully saturated with water.
plotted against saturation for the same rock, there are The bell-shaped curve is characteristic of viscous
peaks of attenuation at low saturation (e.g. 1%) and at damping. After Paffenholz and Burkhardt, 1989;
high saturation (e.g. above 60 to 90%), depending on and Walsh, 1995.
whether loading is in pure compression or in shear,
where the peak attenuation comes only at greater sat-
P and S waves passing through a medium exert oscilla-
uration. It is believed that the attenuation peak at very
tory stresses which can be resolved into normal and shear
low saturations is the result of viscous losses in fluid
components in the plane of each pore space. Attenuation
trapped in microcracks, which are filled first due to
can be demonstrated both for the normal component
stronger capillary forces where apertures are very small.
and for the shear component. The ratio Qp/Qs is 1 for
Over a wide range of saturation from a few percent to
dry rocks and is 1 when almost fully saturated, as we
some 50%, no change of attenuation occurs, but as con-
have seen earlier. It appears from the model of Mavko
tinuous saturated regions arise, fluid pressures rise in
and Nur, 1979, that the state of saturation of the flat
response to the pore volume reduction caused by the har-
cracks (or rock joints) rather than the overall saturation of
monic compressive wave, and energy loss becomes fre-
the rock is the most important factor for the attenuation.
quency dependent due to viscous dissipation. However,
at very high frequencies, no fluid transfer occurs and the
saturated region responds elastically. At very low fre- 10.3 Effect of confining pressure on
quencies, flow occurs, but if viscous stresses are low, dis- seismic Q
sipation of energy may be negligible.
The bell-shaped frequency dependence of seismic Resonant bar techniques for the sonic frequency range,
1
Q occurs at the intermediate frequencies, when vis- and pulse transmission techniques for the ultrasonic
cous dissipation is not negligible. This is illustrated in frequency range were used by Lucet and Zinszner,
Figure 10.14. The narrow peaks of attenuation in both 1992, to demonstrate that not only frequency range, but
shear and hydrostatic compression were suggested by also confinement can affect the seismic quality Qseis. Their
Walsh, 1995, to be the result of the ‘squirt’ phenom- 3 to 7 kHz and 500 kHz testing with some 30 rocks
enon (Mavko and Nur, 1975). that included limestones and sandstones, included con-
At low frequency, pore fluids influence the attenuation finement to 45 MPa and water saturation. Pore pressure
due to their lack of rigidity, compressibility and density, was fixed at 1 atmosphere. Care was taken to select core
while at higher frequencies, attenuation occurs due to from adjacent samples in the same homogeneous quar-
viscous and inertial forces. Nur, 1973, interpreted tem- ried block, so that the effect of different frequencies
poral velocity anomalies as evidence of dilatant strain could be truly compared.
and varying pore water saturation in the crust prior to Figure 10.15a shows a set of results for a sandstone,
certain earthquakes. Mavko and Nur, 1979, showed that in which sonic and ultrasonic attenuation as a function
even a small amount of water can dramatically enhance of increasing confining pressure are (in this case) simi-
the attenuation, when very flat pores (or joints) are pre- lar. The vertical scale of 1000  Q1 shows that Qp (or
sent. This is because high pressure gradients cause (micro) QE), increased from about 6 to nearly 100 as a result of
flow at the contact between wet and dry pore space. confinement. This can be seen by ‘inserting’ seismic Q
Seismic quality Q and attenuation at many scales 191

according to frequency. The ultrasonic Qp1 (attenu-


ation) is significantly higher, or Qp numerically much
smaller (6 to 10) than for the sonic tests, where QE ranges
from 7 to 100 or more, as confinement is increased.
The authors interpreted these differences as being due
to scattering of waves due to ‘density’ heterogeneities in
the case of the limestone. Another limestone which was
fine-grained showed less dramatic separation of behav-
iour as a result of frequency differences, and almost
negligible effect of confining pressure. Seismic Q values
were in this case a more or less constant 50 (ultrasonic)
and a more or less constant 100 (sonic), over the full
confining pressure range.
In an important series of tests on two sandstones,
Prasad and Manghnani, 1997, investigated not only the
effects of effective stress change, but also pore pressure
changes on the P-wave velocity and attenuation Qp1
(a) Their experimental set-up, which is simply and clearly
illustrated, has been reproduced in Figure 10.16. This
figure defines Pc and Pp, and the difference Pd  Pc -
Pp is found in subsequent figures showing their results.
The two sandstones investigated, Berea and Michigan,
had bulk densities of 2.28 and 2.36 gm/cm3, and corres-
ponding porosities of 21.2% and 16.9%, respectively,
causing the higher velocities in the Michigan sandstone.
The Berea sandstone had visible bedding planes and
weakly cemented angular grains with microcracks. We
can therefore select this sandstone for reproducing some
of the author’s important results.
These results, and equivalent ones for the rounded-
grained and less porous Michigan sandstone, enabled
the authors to differentiate the pore pressure depend-
ence of the two sandstones. Referring to the classic
effective stress equation:

Pe  Pc  nPp (10.3)
(b)

Figure 10.15 Sonic (resonant bar extensional mode), and ultrasonic where n is the effective stress coefficient (Biot, 1962,
measurements, a) on a saturated sandstone, and b) on Todd and Simmons, 1972), the authors found that
a saturated crinoidal limestone, as a function of con- both the Berea sandstone and the Michigan sandstone
finement. Lucet and Zinszner, 1992. had values of n that reduced from about 0.78 and 0.62
respectively, when the confining pressure was high.
values down the right-hand axes at convenient arith- These results applied to experiences in interpreting Vp.
metic intervals, giving Qseis values of 5, 10, 20, 50 and In the case of Qp, equivalent results were 1.10 and 0.86,
100. To one with a rock mechanics background, a reducing to 0.81 and 0.71, respectively. In other words,
resemblance to E-modulus increases with confinement Vp and Qp measured at elevated pore pressures and
is seen in both sets of sonic data, with units of GPa. elevated confining pressures are governed by effective
More of this will be seen later. stress coefficients significantly less than the classic n  1
In the case of a crinoidal limestone shown in Figure obtained for more permeable media. The authors
10.15 b, there is clear separation of the attenuation Prasad and Manghnani cited differences in the type of
192 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(a)

(b)
(b)
Figure 10.16 Schematic diagram of the ultrasonic pulse transmis- Figure 10.17 Vp and Qp as a function of effective confining pres-
sion experiments of Prasad and Manghnani, 1997. sure for two sandstones, in this case at a pore pres-
sure of 1 atmosphere. Prasad and Manghnani, 1997.

contact areas between the grains in the two sandstones likeness to seismic Qp was noticed. This can be seen in
as the reason for the differences in the pore pressure broad terms in Figures 10.15, 10.17, 10.18 and 10.19.
dependencies of the two sandstones. The likeness of Emass in GPa and in situ Qp has contin-
Before leaving the above results of confining pressure ued to be seen in field data reviewed.
on Qp (there will be more data in some other chapters), Almost all rock mechanics modulus data, from labora-
it may be of interest to mention a finding, now a con- tory testing representing near-surface to kilometre depths,
viction, first noted when writing the chapter dedicated and from in situ testing at dam sites and deep tunnel
to rock physics results (Chapter 13). It was finally recog- deformation back-analysis, show moduli within the
nised that the variation of Qp with confining pressure extreme range of 1 to 150 GPa, most commonly 5 to
resembled the well known rock mechanics effect of 75 GPa. In exceptionally weathered, weak, or clay-bearing
triaxial confinement on the E-modulus of rock samples. conditions, moduli can reduce to 0.1 GPa, where ‘total’
When the latter is expressed in GPa, quite remarkable attenuation in less than a wave length no doubt occurs.
Seismic quality Q and attenuation at many scales 193

(a) (a)

(b)
(b)
Figure 10.19 Changes in Vp and Qp in Berea sandstone: in both
Figure 10.18 Changes in Vp and Qp in Berea sandstone: in both cases with differential (or effective) pressure Pd con-
cases with pore pressure PP constant. Prasad and stant. Prasad and Manghnani, 1997.
Manghnani, 1997.

1. 1/Qs  1/Qe  1/Qp  1/Qk (for low Vp/Vs with


partial saturation)
10.3.1 The four components of 2. 1/Qs  1/Qe  1/Qp  1/Qk
elastic attenuation 3. 1/Qs  1/Qe  1/Qp  1/Qk (for high Vp/Vs with
full saturation)
Before looking at the (non-linear) effects on attenu-
ation of samples loaded towards fracturing, at the end of The Batzle et al., 2005 forced-deformation apparatus
this section on laboratory tests with confining pressure, was capable of applying frequencies from 0.3 Hz to
it is appropriate to refer to Batzle et al., 2005 laboratory 2,000 Hz, with strain amplitudes below 107. Note
testing of intact, porous samples in a so-called forced- that the latter is very low. Micro-valves were used to
deformation apparatus. These authors’ tests neatly demon- control fluid movement into or out of the samples, in
strated the relative magnitudes of the elastic attenuation response to the dynamic loading. Both brine-saturated
components. and partly brine-saturated states were investigated.
According to Nur and Winkler, 1979, the different Batzle et al., 2005 found that opening or closing
modes of elastic attenuation (1/Qk  bulk, 1/Qp  their sample boundaries to fluid, using special micro-
compressional, 1/Qe  Young’s and 1/Qs  shear) are valves, caused a significant change in the velocity and
related to each other through inequalities. dispersion values, when at full saturation. Two sets of
194 Rock quality, seismic velocity, attenuation and anisotropy

their experimental results are reproduced in Figure across the boundary was absent at high frequencies, due
10.20. The authors noted with the open boundary, that to the lack of time to reach (pressure) equilibrium.
low frequencies caused the rock-fluid conjunct to Beyond 100 Hz, saturated samples (with open bound-
behave as if partially saturated. The fluid movement aries) showed a low cumulative elastic attenuation.
Batzle et al., 2005 cited the fundamental coupling of
attenuation, velocity and frequency, from the illustra-
tive Cole and Cole, 1941 developments in dielectrics,
which were applied to attenuation measurements by
Spencer, 1981. The authors added the effect of fluid
mobility and partial saturation, and indicated the typ-
ical measurement window, in Figure 10.21
The authors also addressed the more complex question
(a) of attenuation and frequency dependence, or dispersion,
in samples of shale, finding that dispersion had strong
directional dependence. Because of the low permeability
and inhibited fluid motion, yet observed dispersive results
because of strong attenuation, they suggested that inter-
actions among clay particles and between the clays and
bound water may be responsible.
They also showed that viscous fluids like heavy oil
had their own internal viscous losses, which could con-
(b) tribute to overall rock attenuations. The conclusion
from their studies of the multiple components of (elas-
Figure 10.20 The relation between the four different modes of
elastic attenuation with a) partial brine saturation,
tic) attenuation was that attenuation-related attributes
where attenuation due to the bulk modulus dom- extracted from seismic data have to take such control-
inates, and b) 100% brine saturation and an open ling parameters into account.
boundary, where at low frequencies 1/QK and 1/Qp Problematic here is that the fracturing or joint set
dominate, as fluid can flow in and out of the developments in a rock mass will often be concentrated in
samples. Batzle et al., 2005. the higher modulus layers, whose internal attenuation

Figure 10.21 The Cole-Cole relation coupling velocity, attenuation and frequency from the field of dielectrics. Cole and Cole, 1941. This
was applied to attenuation in rock by Spencer, 1981. Batzle et al., 2005 also indicated the approximate dispersive effect of low
or high fluid mobility, and of partial saturation.
Seismic quality Q and attenuation at many scales 195

components have thereby changed, or are different, agreement of observations with the intrinsic attenuation
from surrounding rock. Fluids of different viscosity in mechanism of frictional sliding, developed by Walsh,
the differently fractured layers will add to the challenge 1966, and good agreement with the semi-empirical
of inverting data. pressure-dependent theory of Johnston et al., 1979, that
is also based on Walsh, 1966. They also cited studies of
10.3.2 Effect on QP and QS of loading scattering attenuation in micro-fractured marble where
rock samples towards failure the scattering attenuation theories of Hudson, 1981 and
1990 (the first-order scattering model), did not predict
An important contribution was made to our under- sufficient attenuation in relation to test results. Wulff
standing of the influence of fracturing on attenuation, et al., 1999 found that crack density squared was needed
with simultaneous velocity effects, by Wulff et al., in the Hudson, 1981 model, rather than a linear relation
1999. The authors made a careful study of the seismic to crack density, to explain their own results.
effects of microfracturing during constant, low strain The authors’ tests on blocks of tuffaceous sandstone
rate uniaxial compression testing, up to and beyond the and granite, measuring 100  100  250 mm, were
point of microfracturing. They tested tuffaceous sand- conducted under ‘room-dry’ conditions, following four
stone and granite samples, both related with Hot Dry weeks of drying at room temperature. New cracks were
Rock projects in Japan. assumed to be dry ‘or at least not to have absorbed
As the authors pointed out, attenuation was not enough water molecules to permit fluid flow during the
directly related to the strength and elastic moduli, but to time of the experiment’.
mechanisms such as fluid flow, friction and scattering Several of their very interesting results are repro-
due to microcrack and crack density effects. Testing duced here. Figure 10.22 shows the separate effects of
only dry specimens, they concentrated on interpreting axial strain (with associated development of microc-
the relative roles of scattering and friction. racking) on P- and S-wave velocities, and on Qp and
They reviewed several studies of attenuation in dry Qs, for two samples of tuffaceous sandstone (t2-l, and
rock, (slate, sandstone, gabbro), that indicated good t1-f ). The wave propagation was perpendicular to the

(a)

(b)

Figure 10.22 Effects of uniaxial stress-strain: symbol (o), and the associated microcracking, on Vp, Vs, Qp and Qs for two tuffaceous sandstone
samples (100  100  250 mm, n  18.6%,   2.05 gm/cm3) from Japan. Note sample failures at 25 and 39 MPa. a) The
upper pair of results (sample t2-l), have wave propagation (400 kHz) perpendicular to the loading direction, and therefore perpen-
dicular to dominant microcracking. b) The lower pair (sample t1-f), have wave propagation (also 400 kHz), parallel to the loading
direction, causing increased velocity and less attenuation with increased load. (solid symbols: Vp and Qp). Wulff et al., 1999.
196 Rock quality, seismic velocity, attenuation and anisotropy

‘Poisson expansion’ effect. In the case of the tests on


granite shown in Figure 10.23, seismic Q was also meas-
ured perpendicular to loading, so registered similar
reductions to Qp and Qs, following a slight increase in
Qs during the first half of the loading.
The authors investigated theoretical crack densities
(crack number density  crack radius cubed), based on
the theories of Hudson, 1981 and 1990. They inter-
preted a non-linear increase in attenuation with crack
density as being due to pressure-increased crack sizes in
addition to crack density increase. The frequency
(a) dependence of the P-waves, proportional to approx.
f 2 or 3, suggested attenuation by scattering and possibly by
friction. They therefore investigated the (Rayleigh)
scattering attenuation predicted by the Hudson 1981
model, and found that only when using the largest
plausible crack dimensions could they explain the total
attenuation, if scattering alone was responsible.
The attenuation mechanism caused by frictional slid-
ing along the tapered tips of microcracks, according to
Mavko, 1979, was also investigated, giving a good fit to
the total attenuation of two of the four samples, based
on their assumptions. The mechanism is independent
of frequency, which the authors found consistent with
(b) the fact that the measured total attenuation was less fre-
quency dependent than expected if scattering was the
Figure 10.23 Effects of uniaxial stress-strain: symbol (o), and the
only mechanism. Figure 10.24a compares the scatter-
associated microcracking, on Vp, Vs, Qp and Qs for a
ing attenuation calculated with the Hudson, 1981,
granite sample (100  100  250 mm, e.g., average
grain size 1.3 mm, n  1.4%,   2.62 gm/cm3) from
model, based on maximum plausible crack sizes, with
Japan. Note failure at about 150 MPa. Wave propaga- measured data, and Figure 10.24b shows the calculated
tion (400 kHz), is perpendicular to the loading direc- attenuation due to crack-tip friction, following Mavko,
tion, and therefore perpendicular to dominant 1979. The authors concluded that attenuation in the
microcracking, causing the reduced velocities and rocks investigated was probably by a combination of
increased attenuation (reaching a minimum Qp  5). frictional attenuation and scattering.
(solid symbols: Vp and Qp). Wulff et al., 1999. A laboratory study involving ‘flaws’, this time nat-
urally existing, concerning the cavities in carbonate
loading direction in Figure 10.22a, and parallel to the rock, such as vugs or karsts, was described by Hackert
loading direction in Figure 10.22 b. In each case the and Parra, 2003. These cavities cause scattering attenu-
axial load – axial strain curves are given by (o) symbols. ation like the fracturing seen above, but quantification
Equivalent results for a granite specimen, with wave is difficult due to the unknown scale and structure of
propagation perpendicular to the loading direction, are the cavities. The authors described the use of X-ray
shown in Figure 10.23. computerized tomography scans to obtain the exact
The increase of seismic Qp from about 18 to 30 (and vug structure of two cores. They then used 3D finite-
Qs from about 30 to 50) in the case of the measurement difference modelling to determine the P-wave
parallel to the loading direction, are both potentially scattering attenuation at ultrasonic frequencies.
recognisable as ‘deformation modulus’ results, if the lat- Qseis in the saturated states were as low as 8 and 15
ter were expressed in GPa. On the other hand, the meas- near the source frequency of 250 kHz. The two cores had
urements of seismic Q made perpendicular to the respective total porosities of 32.1% and 16.6%, with
loading direction, showing both Qp and Qs reducing CT-computed vuggy porosities of 13.4% and 4.5%. The
from about 18 to 7 or 10 were actually registering a respective dry-state P-wave velocities were 3.97 and
Seismic quality Q and attenuation at many scales 197

Figure 10.25 Basic test set-up, for conducting normal loading and
dynamic testing of joint samples, with add-on facil-
(a) ities for hydraulic testing and contact area estima-
tion using Woods Metal. Pyrak-Nolte et al., 1990.

‘divided’ by single natural joints or fractures. Landmark


work was done in this area by Laura Pyrak-Nolte and
colleagues Neville Cooke and Larry Myer, with import-
ant links to the hydraulics of joints or fractures via Paul
Witherspoon. This pioneering research, originating from
the University of Berkeley and from Lawrence Berkeley
Laboratory, followed on from the rock mechanics devel-
opments of Goodman twenty years previously, and the
theoretical geophysics of Schoenberg, and represents
one of the few and important links between rock
mechanics, hydraulics and geophysics.
(b)
Most of Pyrak-Nolte’s and colleagues’ better known
Figure 10.24 a) Calculated scattering attenuation for sandstone sam- work was focussed on the behaviour of just three sam-
ple t2-l, using the Hudson, 1981 method, assuming ples of joints in quartz monzonite from Stripa Mine
crack sizes of 600
m. Solid lines represent the model ‘granite’ in Sweden. Even the much described sample
with randomly oriented cracks, and dashed lines rep- numbers E30, E32 and E35 are sometimes referred to
resent cracks oriented in the loading direction. b) by geophysicists. These robust samples were subject to
Calculated attenuation due to friction, using the numerous tests, on numerous occasions, and have given
Mavko, 1979 model, based on estimated crack dens- the profession important insight into ‘fully-coupled’
ities for randomly oriented cracks. Wulff et al., 1999. earth science behaviour. We will review different aspects
of this work in this and later chapters.
4.25 km/s, and dry densities 1.85 and 2.20 gm/cm3. The An understanding of the basic principles for their tests
authors observed that if the vugs had been karsts 1000 is given in Figure 10.25. Besides the dynamic testing
times larger (about 5 m), then the attenuation would under normal load, as indicated, there was the possibil-
have been seen at seismic frequencies in the range 100 to ity to measure permeability by linear (sector-to-sector)
500 Hz. flow across the circular joint specimens, which had a
diameter of 52 mm. There was also a facility to inject
non-wetting molten Woods Metal into heated joint
10.4 The effects of single rock samples, which upon cooling, gave a measure of the area
joints on seismic Q of the joint available for flow, at the given normal stress
level. Joint roughness, such as JRC was not described,
We will end this section on laboratory tests concerning but a test result was referred to by Pyrak-Nolte et al.,
seismic Q, with an appropriate transitional stage, namely 1987a, where negligible effect of temperature (95°C) on
the seismic Q behaviour of laboratory samples that are aperture was indicated. (This differs from some other
198 Rock quality, seismic velocity, attenuation and anisotropy

experiences of temperature effects on joint apertures


where there is appreciable roughness, e.g. Barton et al.,
1985, Barton, 1999, Barton and Makurat, 2006).
The hydraulic apertures of the three joints E 32, E 30
and E 35 can be interpreted as reducing from approxi-
mately 3, 5 and 26
m to approximately 1, 1 and
6
m, as the joints were closed by measured amounts of
approximately 4, 9 and 22
m (Pyrak-Nolte et al.,
1987a). Thus E/e was in the low range of 1.1–2.25,
suggesting quite planar joints, in relation to the Barton
et al., 1985 model for the JRC-controlled measured
inequality E  e, or E  e usually seen. A prelim-
inary estimate suggests that JRC may have been in the
(a)
range of only 2 (two cases?) to 4, which would readily
explain the relatively high stiffness and small closures
under stress, exhibited especially by two of the three
Pyrak-Nolte et al., joint samples (E30 and E32).
Figure 10.26a shows joint (or fracture) deformation
versus normal stress for one of a series of load-unload
cycles. The equivalent ‘specific stiffness’ (the inverse of
tangent slopes) for these three load-deformation events,
is shown in Figure 10.26b. These roughly 1,000 to
30,000 MPa/mm normal stiffnesses are of the same
order of magnitude as the results for fresher, i.e. stiffer
joint samples in various hard rock types, tested by
Bandis, 1980. (See Bandis et al., 1983 and joint stiff-
ness data reproduced in Chapter 16).
(b)
Pyrak-Nolte et al., 1990 performed dynamic tests
both across the three joint samples and across intact sam- Figure 10.26 a) One set of load-deformation results for the three
ples taken from adjacent core. (Stripa ‘granite’/quartz quartz monzonite (Stripa granite) joint samples. b)
monzonite, ␥  2.65 gm/cm3, Young’s modulus E  Specific stiffness (inverse tangent slopes) as a function
60 GPa). The three pairs of samples had equal length and of normal stress, for the three joint samples tested dry.
diameter (77  52 mm). Pyrak-Nolte et al., 1987a.
Figure 10.27 shows three sets of ultrasonic (‘0’ to
1.5 MHz) P-wave amplitude spectra, conducted in the spectral amplitudes obtained from a non-attenuating
dry state in this case, for the three pairs of ‘companion’ cylinder of aluminium of identical dimensions, and
samples (jointed, and adjacent intact). The comparison loaded in an identical manner to the same loads.
of ‘intact’ and jointed response gives a very instructive The authors, following Johnston et al., 1979, com-
image of the effect of the more open and deformable E pared the spectral amplitudes of the different samples,
35 joint (top), on P-wave transmission, showing select- starting with the dispersive wave equation:
ive filtering of highest frequency. (Note the lower nor- &x
mal stiffness of E 35 in Figure 10.26b, compared to the A  A0 e 2Qc
(10.4)
very stiff and presumably well interlocked E 32 sam-
ple). Sample E 30, with intermediate stiffness, gives where %  frequency
intermediate response. X  travel path length
This fairly long, but necessary introduction to these Q  inverse of attenuation
important tests, brings us to the subject of seismic Q. c  phase velocity of wave
Pyrak-Nolte et al., 1990 calculated seismic Q by taking A0  amplitude at x  0
the ratio of the spectral amplitudes of the intact and The ratios of the spectral amplitudes of the seismic
jointed ‘companion’ samples, in comparison to the pulses transmitted through the companion rock samples
Seismic quality Q and attenuation at many scales 199

Table 10.3 Seismic Qp and Qs calculated by Pyrak-Nolte et al.,


1990, using equation 10.4, for both jointed and
intact, and dry and saturated states, at two levels of
normal (axial) stress. The results for the most
deformable E 35 joint sample are selected, as the
25
m joint closure with 85 MPa stress increase is
considered realistic for non-planar joints.

Qp Qs
Specimen type
and test condition 2.9 MPa 20 MPa 2.9 MPa 20 MPa

Sample No. E 35 E 35 E 35 E35


Jointed-dry 7 14 12 23
Intact-dry 12 39 32 71
Jointed-wet 9 30 28 39
Intact-wet 15 51 41 56

compared to the aluminium (A/A1), were then calcu-


lated (by Johnston et al., 1979) as:

A fx A
ln   ln 0 (10.5)
A1 Qc A 01

where f  %/2
The authors gave a comparison of seismic Q calcu-
lated from this equation, for both the dry and saturated
states. A selection of their interesting results is repro-
duced in Table 10.3.
The marked reduction in attenuation when loading
the jointed specimen at 20 MPa instead of 2.9 MPa is
typical of in situ response. Qp increases by a factor of
2 when dry, and by a factor of 3 when wet. It is inter-
esting to note however, that the less attenuating intact
specimen shows Qp increasing by a factor of at least 3,
both when dry and when wet. The seismic waves are of
course transmitted perpendicular to the microcracks
most likely to be closed by the axial stress.
Pyrak-Nolte et al., 1990 made an alternative seismic
Q calculation, because of the non-linearity of the spec-
tral ratio data. By assuming that A0/A01  1 in equation
10.5, they were able to re-arrange the equation and
Figure 10.27 Comparison of intact and jointed sample response express seismic Q as a function of frequency. Figure
to ultrasonic P-waves up to 1.5 MHz frequency. 10.28a shows the result of applying equation 10.6 to
The magnitude spectra show the positive, magnitude- the data from the dynamic tests on jointed sample E
increasing, effect of higher normal stress. Sample 30. This joint had an intermediate level of normal stiff-
E 35 has least normal stiffness, due to its 25
m of
ness in relation to E 35, and to the least deformable,
closure under normal stress to 85 MPa. It demon-
stiffest joint E 32 (Figure 10.26).
strates the maximum filtering of higher frequencies,
compared to the high stiffness sample E 32, which
only closed some 5
m under 85 MPa stress. Pyrak- fx
Q  (10.6)
Nolte et al., 1990. clnA/A1
200 Rock quality, seismic velocity, attenuation and anisotropy

(based on the theory of Schoenberg, 1980), with


dynamic normal stiffness varying from 6.4  1012 to
1.6  1013 Pa/m. In rock mechanics units, this is more
easily understood as 6,400–16,000 MPa/mm, in fact
typical for the pseudo-static normal stiffness of fresh
rock joints at high stress levels. (See Chapter 16.) The
theoretical curves show good correspondence to the
interpreted data using the spectral amplitude ratio
method described above.
Before leaving these interesting studies for the time
being, the earlier referred suspicion that deformation
modulus (in GPa) is similar to seismic Q, will be
addressed again. If we ‘insert’ a virtual Qseis scale down
the right-hand axes in Figure 10.28, we obtain ‘simple’
magnitudes for seismic Q of 5, 10, 20, and 100 for
1/Qseis values of 0.2, 0.1, 0.05 and 0.01. The jointed
sample shows the lowest Qseis values when measured at
the ‘lower’ frequencies of 0.2 MHz, and values varying
from 5 to about 20 as normal stress increases from 2.9 to
70 MPa while the solid control sample, at 0.5 MHz
shows Qseis of 8, 15, 20 and 90 as stress was raised from
2.9 to 10, 20 and 70 MPa. These results remain in the
typical range of moduli (when expressed as GPa), but the
low stress value of Qseis seems to be lower than expected.
In-seam seismic measurements in coal have been
used for a number of years to indicate the state of stress
and fluid drainage in this fine-structured, deformable,
Figure 10.28 a) The frequency-dependent attenuation calculated low velocity material, which in some ways resembles
for the medium-stiff sample E 30 (dry), with com-
a miniature (and property-scaled) roughly cubically-
parison to the lesser attenuation of the intact (dry)
jointed rock mass. A set of laboratory test data, includ-
companion sample. Normal stress levels were 2.9, 10
and 70 MPa. Small circles were the result of calculation
ing effects of the dry to fully saturated state, with
using equation 10.5, at a specific frequency of confining pressures from near-surface, up to mine-
0.5 MHz. b) Analytical solutions assuming increasing relevant levels (2 to 40 MPa) was given by Yu et al.,
values of dynamic normal stiffness, as described in the 1993, using a transversely isotropic Permian coal from
text. Pyrak-Nolte et al., 1990. Tower Colliery, Wollongong, Australia.
Some of the key results of these comprehensive studies
Figure 10.28a shows the dispersive results of the seis- are reproduced in Figures 10.29 and 10.30. The strong
mic Q calculation, using this equation for the tests effect of water saturation, which tends to fill the flat (low
conducted on E 30, when in the dry state. The small cir- aspect ratio) cracks and cleats in the coal, is evident in all
cles in the figure were calculated using equation 10.5, at a the data. This miniature ‘rock mass’ also displayed the
specific frequency of 0.5 MHz. The similar shape of the classic anisotropic effects of lower Vp perpendicular to
curves for the jointed and intact samples was interpreted bedding, and higher Vp parallel to bedding, with 45°
by the authors as evidence for similar (closure-under- wave transmission giving intermediate values. S-waves
stress) behaviour of both the joint and the microcracks were little affected by saturation, which is also a tradi-
most likely to be (partly) closed by the axial stress. tional result, when S-wave splitting is not involved. In
The theoretical curves shown in Figure 10.28b were Chapter 15 we shall see that polarized split shear-waves
developed by the authors, by assuming a joint or fracture are affected by degree of saturation and even fluid type,
density of 1 per 77 mm, as tested, in an otherwise non- due to changed joint or fracture compliances.
attenuating medium (clearly a simplification). The single At higher confining pressures, the water content (one
fracture was represented by a displacement discontinuity of the contributing causes of attenuations) was reduced
Seismic quality Q and attenuation at many scales 201

Figure 10.30 a) The spectral amplitude behaviour across and par-


allel to the coal’s bedding, showing the strong effect
Figure 10.29 Vp as a function of confining pressure (pore pres-
of saturation on wave transmission. b) Qp for dry
sure  atmospheric), for dry and saturated macro-
and saturated coal specimens, as a function of con-
bedded coal specimens. The specimens were loaded
fining pressure (pore pressure  atmospheric).
and dynamically tested in three orientations relative
Yu et al., 1993.
to bedding. Yu et al., 1993.

by partial closure of the fine cubic structure of cracks the dynamic Poisson’s ratio (e.g. about 0.4) in relation
and cleats. Vp (dry) and Qp (dry) approached the values to typical intact rocks. The effect of higher stress in the
of Vp (saturated) and Qp (saturated) as confining pres- coal reducing the difference in seismic Q between dry
sure was increased. The seismic Q values were lowest and saturated conditions was assumed to be because of
for the four dry specimens, while there was much less both increased frictional resistance along the cracks,
attenuation for the four saturated samples. The authors and due to the reduced water content caused by the
emphasised the fact that the decrease in attenuation for closing cracks. Again we see the general trend of seismic
fully-saturated specimens of coal, with its low aspect Q increasing with stress in a similar manner to defor-
ratio cracks, differed diametrically from the usual result mation modulus.
for sandstones. This could be questioned based on ear- Thanks to these excellent laboratory Q-studies, the
lier results, but the point is made that cracks close eas- scene is now set for going into the field, to see fractures
ier than equant pore space. and rock joints (not forgetting the ‘ever-present’ micro-
Yu et al., 1993, also emphasised the relatively low cracks), in their in situ seismic Q environment. First we
values of dynamic E-moduli for the coal (e.g. about will look at some near-surface seismic Q, including
8 GPa: at low stress?) and the relatively high values of some quite shallow studies in reservoir-type sediments.
202 Rock quality, seismic velocity, attenuation and anisotropy

This will be followed by seismic Q at great depth con- When moving to in situ scale, joint spacing (captured
nected with earthquakes and continental-scale studies. in RQD) and the number of joint sets (Jn) are clearly
Finally we will return to ‘medium’ petroleum reservoir going to have an influence on scattering losses, just as
depths at the end of this chapter on seismic Q, where microcracks and induced rock sample cracking and
the potential economic rewards of understanding Qseis individual joints are seen to influence scattering and
are pressing further developments. Understanding intrinsic losses in laboratory samples, as shown by
petroleum reservoir behaviour, first from more rock Wulff et al., 1999, and Pyrak-Nolte et al., 1990.
physics (laboratory) contributions, then from in situ It is also intuitively reasonable to suspect that the con-
anisotropy effects, (i.e. shear wave splitting), forms dition of joints – their degree of interlock as determined
much of the material in Chapters 13 to 15. both by stress level and by roughness (Jr), and the pres-
ence or absence of mineral coatings or clay fillings (Ja) –
will have potential influence on (micro) permeability
10.5 Attenuation and seismic Q from and therefore on potential squirt losses, when there is a
near-surface measurements variable degree of saturation, as is frequently the case in
the near-surface.
Seismic reflection and refraction techniques used to illu- Joint characteristics are also expected to influence
minate major features beneath the earth’s surface obvi- eventual frictional losses, if magnitudes of continuum-
ously depend on the existence of seismic wavelets. based shear strain (actually discrete micro-displacement
There may not always exist sufficient seismic imped- discontinuities), are of sufficient magnitude. The fact
ance contrasts between rock boundaries to cause reflec- that seismically determined in situ dynamic rock joint
tion or refraction. But absorption is continuous, and compliances, the inverse of stiffnesses, have immediately
significant information can accumulate on the pro- recognisable (i.e. expected) magnitudes, and are
gressing wavelets. Ecevitoglu and Bingol, 1999 pointed expressed in the same (but inverted) units as in the rock
out that the absorption information may be crucial as mechanics of discontinua (e.g. Bandis et al., 1981,
far as the rock’s consolidation, porosity, fractures, and 1983, Pyrak Nolte et al., 1990), is further justification
fluid contents are concerned. Near surface measure- for looking also beyond microcrack-scale, for the contri-
ments of seismic Q are however complicated by the butions of the jointing of the rock mass, to both scatter-
presence of near-surface weathering layers, and of ing and intrinsic losses.
course by faulting. The seismic energy will be strongly In the last section on laboratory testing of the
attenuated, and waveforms may also be distorted. dynamic response of joints under load, the results of
Pyrak-Nolte et al., 1987a and 1990, showed ‘static’ nor-
mal stiffnesses for the most deformable joint (E35, Figure
10.5.1 Potential links to rock mass 10.26b) that varied from 5,000 MPa/mm at 5 MPa nor-
quality parameters in mal stress, to about 30,000 MPa/mm at 70 MPa normal
jointed rock stress, each in the dry state. These authors’ ultrasonic
joint measurements, showed for the same sample E35,
With our progression from laboratory to field scale, a dynamic normal stiffness varying from ‘only’ 4,000 to
implicit links between Qseis and rock mass quality Q 7,600 MPa/mm (dry), and from 9,500 to 15,000
(Barton et al., 1974, see Appendix A), can apparently MPa/mm (saturated), at comparable low normal
be seen, due to the logical results of near-surface, seis- stresses of 2.9 and 6.0 MPa respectively. At normal
mic Q that rapidly increase with depth. This also mir- stress levels of 70 MPa, the authors showed dynamic
rors the way that rock mass deformation modulus, normal stiffnesses of 32,000 MPa/mm when dry, and
calculated from the rock mass Q-value, also increases 59,000 MPa/mm when saturated.
with depth, to match in situ measurements of this fun- Only the latter is higher than the ‘static’ stiffness,
damental rock mass parameter. This seems to be showing similar relative increases as the Edynamic to
because the features of a rock mass that are described by Estatic inequality that reduces strongly with high quality
a rock mass characterization method, such as rock mass rock joints and rock masses.
quality Q, contains elements of the in situ medium Of importance for in situ uses of seismic Q, the
deemed potentially important for both the intrinsic authors Pyrak-Nolte et al., 1987b noted that velocities
and scattering attenuation mechanisms. soon reached the level of the intact rock, when using
Seismic quality Q and attenuation at many scales 203

high frequencies. They showed that this ‘seismic clo- quality value for a good quality but jointed crystalline
sure’ could occur at lower stress levels when the joints bedrock with two to three joint sets. It might have
were less stiff (as for their sample E 35, Figure 10.26b), Q-parameters as follows (see Appendix A).
and at higher stress when the joints were stiffer. Most
importantly, even when the effect of a joint on velocity 100 2 0.66
was almost ‘erased’ by stress and high frequencies, the Q     20
6 1 1
change of amplitude of the transmitted wave, i.e. the
spectral amplitude basis for seismic Q-estimation,
remained very strong. Greater frequency of jointing in a fracture zone
Kang and McMechan, 1994 showed near surface data would probably reduce this value to 1 or less. This is
from N. Texas where the smallest Qp value of 36, and consistent with independent Q-logging results at SKB’s
the smallest Qs of 23, were relevant to the highly vari- Swedish nuclear waste investigation sites, performed on
able surface weathered zone. According to these authors, 4,000 m of core by the writer in 2003.
very few in situ measurements of scattering in the upper Shaw et al., 2004, reported near-zero offset VSP
few metres to tens of metres were available at that time. investigation of Qseis in a 50 to 600 m deep section of a
‘The near surface velocity/density structure may be well through a Faroe Islands Upper and Middle basalt
more variable than previously thought: some of this series, typical of other North Atlantic basalt forma-
variability may have been hidden in previous measure- tions. The source used was a 150 cubic inch air gun
ment that did not explicitly separate intrinsic and scat- fired in a pit under 2 m of water. The receiver was a
tering effects.’ This of course is supported, implicitly, clamped, three-component geophone, with spatial
by engineering experiences: see for example Chapter 1 intervals of 10 m. The authors were able to assess the
and the rapidly changing (laterally and with shallow errors in the Qseis assessment, by testing with slightly
depth) refraction seismic velocities of Sjøgren et al., different receiver separations of 280, 290, 300, 310 and
1979. Such would also imply rapidly changing rock 320 m. Their results, expressed as Qseis versus mid-
mass qualities, deformation moduli, and by implica- point depths from 200 to 450 m, showed Qseis increas-
tion, seismic Q, since in the near-surface, Vp  ing rapidly from about 10 to 50 in the upper third of
3.5  log Q km/s, where Q is in this case the rock mass the well, and levelling off at about 60 at greater depths.
quality of Barton et al. 1974. (See Chapter 5 for rock The effect on seismic Q of lower stress in the upper
quality and velocity variation at shallow depth). levels of the basalts is implied in these and other stud-
A hydraulically conductive, gently dipping fracture ies. The results are also typical of rock mass deforma-
zone at SKB’s study site at Finnsjön, north of tion modulus variation with depth, where rock quality
Stockholm, was imaged using the seismic reflection Q might typically vary from about 2 to 20, based on
method. Amplitude decay curves as a function of dis- Q-logging of numerous basalts.
tance, given by Juhlin, 1995, showed that a seismic Payne et al., 2005 described a seismic (sparker
quality Qseis of 10 fitted the data, assuming an average P-wave) experiment at a shallow borehole test site in
frequency content of 150 Hz and a P-wave velocity of N.E. England in variously jointed Cretaceous chalk.
5.5 km/s. The value of Qseis  10 was assumed to be Cross-well seismic was performed between three wells,
relevant to the upper 100 metres of this granodioritic at a frequency band width of 500 to 3,000 Hz. Spectral
rock. Juhlin, 1995, considered the result to be consist- modelling was performed to provide Qseis estimates for
ent with higher Qseis values of 30 and 50 at depths of a shallow 30–36 m deep highly jointed zone, with per-
between 200–1100 m in crystalline rocks of compara- meability of about 1 darcy (105m/s), and for a deeper
ble character. Again note the similarity to rock mass (36–50 m), less jointed interval, which had an implied
deformation modulus expectations, when the latter is permeability close to that of the matrix of about 1 mil-
expressed in GPa. lidarcy, or 108m/s. The respective Qp values were
In relation to the Q-value of rock quality, a P-wave 20 and 60.
velocity of 5.5 km/s at 100 m depth in a crystalline, To help assess whether the higher attenuation in the
hard, low porosity rock (Figure 5.36, Part I) suggests a highly jointed zone was mostly caused by scattering
Qc value of about 40, and when compression strength rather than by intrinsic mechanisms, the authors used a
of say 200 MPa are allowed for, the rock mass quality discrete particle numerical model, as described by
Q-value would be about 20. This is a very typical rock Toomey and Bean, 2000. (Although several numerical
204 Rock quality, seismic velocity, attenuation and anisotropy

(This was perhaps surprising, as dynamic stiffnesses


as low as 250 MPa/mm do not imply either high stress
or non-attenuating conditions. Perhaps the discontinu-
ous nature of the modelled ‘fracturing’ had something
to do with this ‘close to infinity’ seismic Q).
When on the other hand, the fracture compliance was
increased by an order of magnitude to 4  101 1 m/Pa
(or when the dynamic Kn was as small as 25 MPa/mm),
a marked reduction in wave amplitude was registered
(see third, attenuated curve in Figure 10.31b). An
apparent seismic Q of 15 was calculated for this attenu-
ating case. Scattering within the highly jointed zone was
(a)
therefore considered a plausible mechanism, presum-
ably because a dynamic Kn as low as 25 MPa/mm (and
its inverted compliance value) were thought to be rea-
sonably representative of the shallow joints or fractures
in the chalk.
As will be shown in Chapter 16, such low values of
normal stiffness indeed imply very low normal stress,
quite consistent with the ‘macro-deformation’ and
comparatively ‘static’ normal closure testing results for
natural, fresh or partly weathered joints reported by
Bandis, 1980 and Bandis et al., 1983. ‘Static’ (extreme
low frequency) testing, perhaps at a rate of only 103 Hz,
(b) displays strong non-linearity in relation to normal stress
level, when loading is from zero to about 60% of the
Figure 10.31 a) Numerical simulation, based on Toomey and
joint wall strength JCS – usually represented by many
Bean, 2000 model, of a random, vertically fractured
section, with source and receiver at opposite sides,
tens of MPa of normal stress. The so-called ‘initial nor-
representing part of the cross-hole experiment in mal stiffness’ described by Bandis, has values of the
jointed chalk. b) Amplitude of received waves, as same low magnitude as assumed above for the dynamic
affected by fractures with two different compliance modelling.
assumptions. Reference case without fractures is the An interesting example of low, near-surface seismic Q
solid line. Payne et al., 2005. structure, with comparison to refraction seismic struc-
ture, was given by Ecevitoglu and Bingol, 1999. They
models will be described in Chapter 15, we may exam- introduced a new methodology to rapidly compute and
ine this model here, in the context of seismic Q). graphically map seismic Q, arguing that absorption
A horizontal compressional wave was applied from measurements are tedious, subject to noise, and not
one side of a random generation of vertical fractures common in everyday geophysics in the near-surface.
having compliant bonds, with the receiver at the other It appears from a necessarily brief (extended abstract)
side. The compressional wave amplitude at the receiver description that the direct wave and all the refracted
is shown in Figure 10.31, and indicates three amplitude- waves were each considered as the first breaks.
time curves. Figure 10.32 shows an example of their tomography-
When the (dynamic) compliance ZN was as large like seismic Q distribution from the near-surface to
as 4  101 2 m/Pa (or dynamic Kn as small as 50 m depth. The high Q anomaly of 20, at 45 m depth,
250 MPa/mm), there was merely a small time shift of was found independently in a conventional seismic
the signal in relation to the homogeneous unfractured refraction interpretation with an upper layer of Vp
case (see dotted, parallel curve). The authors indicated of 1.97 km/s, and a second layer of 4.26 km/s, result-
that this implied that the apparent seismic Q of this ing in a 45 m depth for the refractor. ‘The exact loca-
numerical assembly of particles-containing fractures tion we have found independently from seismic Q
was then ‘close to infinite’. imaging’.
Seismic quality Q and attenuation at many scales 205

Figure 10.32 Seismic Q imaging of an anomaly at 45 m depth with correspondence to a Vp of 4.26 km/s from independent refraction seismic
imaging of this second layer of higher velocity. Ecevitoglu and Bingol, 1999.

10.5.2 Effects of unconsolidated is less easy to accelerate the pore fluid along the pores
sediments on seismic Q (due to inertial forces) than to compress the pore fluid,
as in the flat-pores model of Mavko and Nur, 1979.
Extremely low values of Qp in unconsolidated sedi- Jeng et al., 1999 used artificial source and receiver
ments such as a value of 4 between 60 and 100 m depth pairs, and a frequency-dependent Q estimation, in con-
in sands and gravels (Gibbs and Roth, 1989), and values trast to the conventional spectral ratios with constant
between 2 and 6 for the case of artificial, glycerol- Q assumption. When examining the triaxial-geophone
saturated, random packs of glass beads and coarse sands data with varied (2 m interval) offsets, the frequency-
(Molyneux and Schmitt, 2000), emphasise the character dependent and frequency-independent assumptions
of these unconsolidated and unlithified low Qseis media. reportedly gave ‘dramatic variation’ of Q.
If energy dissipation is small, the seismic quality Qseis The authors carried out experiments at three differ-
(also called the internal friction or dissipation factor) ent sites in Taiwan, but concentrated their attention at
was previously defined as: the Yuan-Lin site in the foothills of central Taiwan,
where two different sources were available. The surface
2 E of the site had a 2 to 3 m thick layer of alluvium and
 (10.7)
Q seis E unconsolidated sediments, overlying a 200 m thick
gravel formation. Their data showed Qp values linearly
where E is the elastic energy stored at maximum stress increasing from between 1 and 3, to between 10 and
and strain and E is the energy loss per harmonic exci- 16, as frequency was increased from 50 to 300 Hz.
tation cycle. Qseis can however apparently be smaller There was marked instability, and therefore lack of lin-
than 2 (i.e. E  E), but alternative definitions of earity, at frequencies beyond 300 Hz. The frequency
Qseis seem to be needed if larger dissipation (i.e. excep- components for the power law Q  kf n were 1.11 and
tionally low Qseis) is measured or assumed. 0.93 for the P and S waves, respectively.
In the case of shallow seismic investigations in sedi- Their investigation using the conventional frequency-
ments, it is likely that Qseis has a frequency-dependent independent assumption for Q, and geophone intervals
component because near-surface layers of sediment tend of 5, 7.5, 10, 12.5, 15 and 17.5 m gave average Q tend-
to be unconsolidated and may contain fluid. This was ing to increase from about 10 to 13 over this range of
verified by Jeng et al., 1999, who measured Qs values as geophone intervals. The modified frequency-dependent
low as 2 to 5 using different sources of energy, and approach at the same location, gave Q values varying
found these lowest values corresponded to the lowest approximately linearly from about 2 to about 18,
frequencies used of about 50 Hz. A roughly 5 times as frequency was increased from 57.5 to 575 Hz.
higher frequency (250 Hz) resulted in about 6 to 8 times The authors conclude that for weathered loose layers
higher Qs-values (16 to 30 approx.). Less attenuation Qseis smaller than 2 is obviously possible, despite the
(higher Qseis) is observed at higher frequencies because it classic formulation of energy loss per harmonic cycle,
206 Rock quality, seismic velocity, attenuation and anisotropy

Figure 10.33 Downhole logs of Vp, density and Qp for ooze and transition to increasing layers of chalk, in sub-ocean studies made during
the Ocean Drilling Program, at the Ontong Java Platform, in the western Pacific. Frazer et al., 1997.

compared to the elastic energy stored at maximum almost linearly from 1.8 km/s at 200 m below sea level,
stress and strain. to 2.6 km/s at 700 mbsl. The reduced density appeared
The authors Frazer et al., 1997, working in the Ocean to be at the base of the ooze-to-chalk transition, where
Drilling Program at the sub-ocean Ontong-Java Platform more chalk was present. In the transition zone the chalk
carbonate sections in the western Pacific, emphasised first appeared in distinct, several centimetre thick
that attenuation measurements made at ultrasonic layers, separated by ooze layers. The thickness and pro-
frequencies in the laboratory, often on disturbed sedi- portion of the chalk increased towards the base, where
ment samples, or estimated from seismic experiments there was more calcium carbonate cement.
over long wave lengths, may reveal little about the The seismic quality Qp for the same hole is also shown
geologic/depth evolution of sediment attenuation. Com- in Figure 10.33. In the loose, high-porosity sediments
paction of loose grains, through diagenesis, to sedimen- the attenuation was assumed to be mostly due to fluid
tary rocks at depth, can be a fragile environment to motion relative to the framework of loosely packed
sample, especially when the shear-wave velocity is lower grains. With greater depth of burial, the number of
than the borehole fluid velocity. points of contacts and their load increases, and friction
The authors used the Schlumberger long-spaced sonic was assumed to become a more important mechanism of
(LSS) tool which has two sources and two receivers in a attenuation. The authors showed porosities as high as
special arrangement. Every 6 inches (0.15 m), four 60 to 70% in the ooze, from 60 to 75% in the transition,
microseismograms were recorded, with three source- and thereafter reducing in the chalk from 60 to 45%.
receiver spacings. The frequencies involved were about Presumably the effects of layering/bedding and perhaps
5 to 25 kHz, lying between laboratory and seismic jointing through the thin chalk beds (?), and related fluid
measurements. flows, contributed to the higher attenuation at the base of
The actual Vp-density log for one of the holes inves- the transition, where Qp was as low as 20. An increase in
tigated is shown in Figure 10.33. The velocity increased Qp was seen where harder chert occurred at greater depth.
Seismic quality Q and attenuation at many scales 207

Figure 10.34 Laboratory tests, shown by stars, and sonic borehole logs of Vp and Qp showing the marked effect of frequency on Qp. Sams
et al., 1997. The sonic logging gave the lowest estimates of seismic Q of all the methods investigated.

10.5.3 Influence of frequency and intrinsic Q contributions are separable by assum-


variations on attenuation in ing that they have different frequency dependencies, as
jointed and bedded rock we shall see in particular from Chapter 13.
A local (micro) fluid flow mechanism is found to be the
When a dynamic load is applied to a rock at low frequen- only mechanism that can account for widely observed
cies, the fluid in compliant large aspect ratio cracks will variations of compressional and shear wave attenuation
tend to be squeezed into the pores and cracks that are less with frequency, both in partially saturated and fully satu-
compliant. As we have seen, the geophysics profession has rated rocks. However, evidence for frequency-dependent
termed this mechanism ‘squirt flow’ (e.g. Mavko and Nur, attenuation from field experiments is apparently less con-
1975; Palmer and Traviolia, 1980; Jones, 1986; Dvorkin clusive: a few cases were reviewed by Sams et al., 1997.
et al., 1995). At higher frequencies inertial effects cause Sams et al., 1997, made a very important contribution
the fluid in the compliant cracks to be less mobile, and in this area of frequency effects, by investigating a
there is lower attenuation, making seismic Q higher. sequence of saturated sedimentary rocks (a finely layered
Laboratory tests which offer the flexibility of using dif- sequence of limestones, sandstones, siltstones and mud-
ferent saturating fluids having different viscosities, in fact stones) using four boreholes drilled to 250–280 m depth
show that there is a peak attenuation when the product of at the Imperial College test site in NE England. They
frequency and viscosity is between 1 and 10 (units of acquired many data sets at widely different frequencies:
Hz.Pa.s) (e.g. O’Connel and Budiansky, 1977). It appears
● VSP experiments (30–280 Hz)
that cracks or joints with aspect ratios of about 103 to
● Cross-hole experiments (200–2300 Hz)
104 cause most of the attenuation (Jones, 1986).
● Sonic logging (8–24 kHz)
The assumption of frequency-independent intrinsic
● Laboratory measurements (0.3–0.9 MHz)
Q and frequency-dependent scattering Q implies that
when the total Q for S-waves (Qs) is smaller than for P-wave velocities for core samples at equivalent depths,
P-waves (Qp), the intrinsic Q is dominant; when it is and ultrasonic Qp estimates measured on core samples
larger (Qs  Qp), scattering Q is dominant over intrin- (each shown by stars) are compared with the sonic log
sic Q (Kang and McMechan, 1994). In fact, scattering results in Figure 10.34. The good correspondence in the
208 Rock quality, seismic velocity, attenuation and anisotropy

Table 10.4 Frequency dependence of Vp and Qp (Sams et al., 1997).

Type Freq. range Median Vp 1000/Qp i.e. Qp

Core 500–900 kHz 3.95 37.0 27.0


Sonic 8–24 kHz 3.48 96.5 10.4
VSP 30–280 Hz 3.20 32.0 31.3
Cross-hole 200–2300 Hz – 63.6 15.7

case of Vp, and the poor correspondence in the case of


Qp at these two different frequencies, is readily seen.
The dependence of median Vp and median Qp meas-
urements on frequency is shown in Table 10.4.
These contrasting results for the frequency depend-
ence of Vp and attenuation (expressed as 1000/Q), are
plotted in Figure 10.35 with curve fitting based on the
authors’ modelling, assuming squirt flow, following
Jones, 1986. This modelling suggests that different sizes
of ‘soft pore space’, i.e. bedding and intra-bedding
joints are sensed by the different frequencies, as a single
aspect ratio fits lower frequency VSP and cross-hole
data, but not necessarily the sonic data.
Sams et al., 1997 estimates of the amount of stiff
porosity (i.e. conventional matrix pores) apparently con-
firmed the results of Mavko and Jizba, 1991, and showed
that the soft porosity responsible for squirt-flow attenu-
ation occupied only a fraction of the total pore space.
Sams et al., 1997, found that the dominant aspect
ratios causing most attenuation were 8.1 to 8.8  104,
even though a much wider distribution of crack (and
joint) geometries is obviously present in this and other
rock masses. We can estimate as an example that if the
mean intra-bed joint apertures were 0.05 mm, the
implied lengths of these would be only about 6 cm.
Perhaps bed-limited jointing, specifically that under least
effective normal stress, could be responsible for most of Figure 10.35 The dependence of P-wave attenuation, Vp and Vs
the attenuation, in view of these relatively ‘large’ aspect on the frequency of measurement in a finely layered
ratios. Sams et al., concluded that the marked frequency sequence of limestones, sandstones, siltstones and
dependence demonstrated by their measurements ‘points mudstones. Sams et al., 1997. The curves relate to
to the amount of information about the rocks that we the author’s modelling of squirt flow losses, using
should be able to obtain from broad-band seismology once the model of Jones, 1986. A range of aspect ratios
we have fully understood the processes that are operating’. was used to represent the ‘soft pores’, or assumed
cracks, intra-bed jointing, and bedding planes.
It is unfortunate that readily obtainable rock quality
descriptions such as RQD and the Q-value are not pre-
sented together with these important geophysics results. Concerning the four sets of Sams et al., 1997, field
If the physical components of these rock masses were data, presenting 1/Q versus frequency for VSP, cross-well,
logged in the conventional manner of engineering geol- sonic and ultrasonic (shown in Figure 10.35), Vogelaar
ogists, possibly it would be easier to understand the and Smeulders, 2005 recently showed that the levels of
potential roles of scattering and intrinsic attenuation in attenuation measured in these field experiments in the
these results. relatively shallow experimental borehole site, exceeded
Seismic quality Q and attenuation at many scales 209

by far, the theoretical prediction of Biot, 1956a,b, com-


paring just the viscosity-based damping of the Biot the-
ory. This comparison is shown in Figure 10.36.
In efforts to improve the fit of theoretical approaches to
the four sets of attenuation-frequency data presented by
Sams et al., 1997, Vogelaar and Smeulders, 2005 mod-
elled a periodically layered porous medium, where the
repeating layers 1 and 2 had pore fluids with different
properties. They applied Biot’s poroelasticity equations,
together with elements of White et al., 1975, who sepa-
rated strain due to the fast compressional wave and strain
due to the slow compressional wave, the former obeying
the wave equation, and the latter the diffusion equation.
They showed that the resulting numerical model
solution based on White’s local flow model, demon-
strated an attenuation rising to 0.05 (Q  20), more Figure 10.36 Comparison of Sams et al., 1997 field data of atten-
than an order of magnitude higher than the Biot the- uation versus frequency, with modelled data from
ory, at a frequency of between 20 and 100 Hz. They the viscosity-based damping of the Biot theory.
Vogelaar and Smeulders, 2005.
stated that an extension of the White model to higher
frequency, made it capable of predicting the levels of
attenuation seen in the field data. The three factors they
investigated showed firstly a maximum attenuation at recorded at a certain distance from an earthquake epi-
a specific frequency, secondly the maximum attenua- centre. (Aki and Chouet, 1975). Seismic coda waves
tion occurred at some specific percentage of gas, and of local earthquakes appear to be produced by back-
thirdly that increasing the gas fraction caused the atten- scattering of waves from numerous randomly distrib-
uation peak to shift towards higher frequencies. uted heterogeneities. The longer the waves travel, the
greater the variety of heterogeneities they encounter.
The later portions of a seismogram may therefore be the
10.6 Attenuation in the crust as result of some kind of averaging of many samples of the
interpreted from earthquake heterogeneities of the intervening crust (Aki, 1969).
coda The spectral contents of the early part of a local
earthquake seismogram depend strongly on the travel
Since the 1960s, the seismic quality of the crust as distance and on the nature of the wave path to the
interpreted from attenuation of earthquake waves has recording station. The coda excitation also depends on
been the focus of much attention. This research was the local geology of the station site, and can be 5 to 8
guided by attempts to find reliable ways of interpreting times larger on sediment than on granite (Aki, 1969).
the precursors of earthquakes. The source, in place of Most coda measurements are made in the 20 to 200
surface-explosives or borehole piezoelectric devices, was seconds time window. As we shall see later, there are
the earthquakes themselves, and their after-shocks. We obvious advantages of in-borehole seismometer loca-
will trace some of the earlier measurements, and tion, at kilometre depths, to help minimize ‘site effects’.
progress from surface recordings to some of today’s The attenuation related to the rate of decay of the coda
down-hole recording of earthquake sources. is termed Qc1 in the geophysics literature, so seismic Qc
has by chance, an identical symbol to rock quality Qc
(Barton, 1995), that is used to describe rock mass quality
10.6.1 Coda QC from earthquake Q (from Barton et al., 1974). This original, widely used
sources and its relation to term for rock mass quality, was at this time also normal-
rock quality QC ized by uniaxial compressive strengths greater or lesser
than 100 MPa, to the form Qc  Q  c/100. This was
Coda waves are the tail of a seismogram (after the arrival done to improve fit to velocity and modulus of deform-
of major wave types such as P, S and surface waves) ation data. As we have noticed, there appear to be some
210 Rock quality, seismic velocity, attenuation and anisotropy

numeric similarities between seismic Qp at shallow depth or almost absent. On the other hand there is no possi-
(i.e. 1 km), and the estimated deformation modulus bility of ever acquiring a reliable measure of deform-
(Emass, or symbol M), when expressed in GPa. ation modulus at extreme depth, without compressing
It may be of interest to observe that the inverse of ultra-small, unjointed laboratory samples, as done by
rock mass quality (Qc1) is roughly proportional to the tectonophysicists in the past. Perhaps, unknown to the
rock mass permeability or Lugeon value for central writer, the E moduli at 20 km depth or 300 (to 500)
ranges of rock quality, when without the complication MPa effective (to total) confining stress, could reach
of clay-sealing of joints that is common near the sur- much higher values than the most typical 50 to 75 GPa
face. (The Lugeon value L  1/Qc, where 1 Lugeon  seen at an order of magnitude smaller depths, for the
107 m/s, Barton, 1999, 2002). By implication, less case of hard crystalline rocks. As we shall see in Chapter
attenuation (high coda Qc) would correspond not only 16, the differential stress ( 1  3) tolerated by small
to high rock qualities Qc, but also to lower permeabil- rock samples can be increased by a factor of 5 to 10, by
ity, and higher deformation modulus. similar magnitudes of confining stress to the above.
(See Figures 16.57a and b).
Aki’s and Chouet’s observations were interpreted as
10.6.2 Frequency dependence of showing the combined effect of variation of coda Qc with
coda QC due to depth effects depth and the frequency-dependent composition of coda
waves. The average coda Qc over the depth range 0 to
The coda Qc was found to increase with frequency, 12 km was 300 in the case of the Stone Canyon site,
though according to Aki and Chouet, 1975, this did some 15 km to the north of the San Andreas Fault, where
not necessarily mean that the coda Qc of crustal mater- average magnitude 1 earthquakes were analysed by these
ial was frequency dependent. The above frequency authors.
effect was thought to be due to the dependence of coda Figure 10.37 shows the regional variations of the
Qc on depth, since waves were scattered from different coda attenuation Qc1 in the range of frequencies from
parts of the earth’s crust. As the primary waves from an 1 to 20 Hz derived from the California Stone Canyon
earthquake spread out, they leave behind a pool of scat- earthquake events, and from the Japanese Tsukuba
tered energy which quickly becomes homogeneous inside Oishiyama earthquake events. Attenuation Qc1 reduces
the pool because of high diffusivity. and coda Qc increases with increasing frequency. At
Since a large volume surrounding the earthquake Tsukuba, where earthquakes are deeper, attenuation
source is ‘sampled’, the seismic coda Qc has been con- Qc1 was lower and coda Qc therefore higher, especially
sidered as a potential measure of the assumed changing at the highest frequency of 24 Hz. (Qc1  0.001,
rock properties, due to accumulation of stress and strain Qc  1000).
in the hypocentral zone (e.g. Chen and Long, 2000). Carpenter and Sanford, 1985, used spectra from 130
Temporal variations of coda Qc before or following digitally recorded micro-earthquakes (M  0.9 to 0.3)
earthquakes have been reported in some cases, while to compute the apparent seismic Q for upper crustal
unfortunately in other cases, no changes have been rocks near Socorro, in the Rio Grande Rift, New Mexico.
noted. Several of these case records will be reviewed in Most of the seismic wave attenuation due to intrinsic
this section of Chapter 10. The apparent frequency absorption and scattering was computed over the fre-
dependence of coda Qc waves can be explained if the quency range of 3 to 30 Hz. Their apparent seismic
coda waves at 1 Hz are primarily composed of surface Q values were found to increase with event distance, for
waves scattered from shallow heterogeneities, while the eight recording stations used in the study. This
coda waves at 20 Hz are primarily back scattering body increase was modelled with a varying thickness, low
waves from deeper heterogeneities in the high Qseis litho- seismic Q, low-velocity layer, lying above a relatively
sphere. At the two sites investigated by Aki and Chouet high seismic Q, high-velocity half space.
(western California and Japan), the coda Qc ranged As illustrated in Figure 10.38, the waves from the
from 50 to 200 at 1 Hz in the shallowest crust (resem- more distant earthquakes would have a greater fraction
bling possible deformation moduli in GPa), to about of their total ray path in the deeper rocks, therefore
1000 to 2000 at 20 Hz in the deeper crust. indicating that seismic Q was greater at depth. The
The latter, by chance, resembles the rock mass Qc seismic Qp and Qs values were found to be less than 50
value itself, assuming that jointing is effectively closed directly beneath the sites at 0.3 to 2 km. This again
Seismic quality Q and attenuation at many scales 211

resembles rock mass deformation moduli magnitudes,


expressed in GPa.
A stronger seismic Q gradient near the surface was
accounted for by lumping all of the low seismic Q mate-
rial into this one layer. Significantly, the near-surface S-
wave Qs values for stations resting on apparently
competent Pre-Cambrian and Paleozoic rock were quite
low (25) and generally less than for stations resting
on Tertiary tuffs. The authors interpreted this as being
due to a greater incidence of open and water-filled frac-
tures in the otherwise more competent rock. Such
would also be consistent with lower rock quality Q-val-
ues, due in particular to higher, near-surface values of
SRF, causing lower Q-values and lower rock mass
deformation moduli. (See Appendix A for Q-parameter
ratings for describing different rock mass conditions).
Qp/Qs ratios ranged from 0.34 to 1.39, and a decrease
of this ratio was generally measured with increasing dis-
tance. Carpenter and Sanford took this to imply vary-
ing degrees of saturation in the upper crustal rocks.
Near the surface, fully saturated rocks have Qp  Qs,
while at depth, partially saturated or dry rocks may
have Qs  Qp. (Winkler and Nur, 1982).
In a gas-producing region of Uzbekistan in 1.4 km
thick Tertiary sediments, Clouser and Langston, 1991,
determined values of Qp (10 to 70) and Qs (10 to 25)
based on a spectral ratio method of analysing after-
Figure 10.37 Regional variations of Qc1 for the frequency range
shocks from the 1984 Gazil earthquake. There was con-
1 to 20 Hz, derived from analysis of earthquake coda jecture that these 10 to 20 km deep thrust-faulting
waves. Aki and Chouet, 1975. (A potential seismic events could have been induced by gas extraction.
Qc scale on the right side of the figure would range Clouser and Langston, 1991, investigated various Qp-
from a minimum of 67 at about 1 Hz, through 100, Qs relations, comparing some theoretical straight line
200 and 1000 at 25 Hz). relations with some in situ measurements. Figure 10.39
shows the stratigraphic section through this 1.4 km of
sediments, and increasing P and S wave velocities
down to the basement. The Qp and Qs relations are
shown in Figure 10.40, together with black dots repre-
senting other authors’ data for comparable sedimentary
rock and rock sequences. The various initials against the
Qp-Qs curves are from their six monitoring stations SP-1,
GSN, OFT, GAZ, TSV and K31. Intuition would now
suggest that these lower Qp and Qs magnitudes, varying
as they do from about 5 to 70, may represent the lower
frequency, nearer-surface sampling of the jointed crust.
Clouser and Langston, 1991, evaluated the following
two equations:

4  V 2
Qs   s Q (10.8)
Figure 10.38 The two-layer model used to interpret event-distance
3 V  p
effects on seismic Q. Carpenter and Sandford, 1985.  p 
212 Rock quality, seismic velocity, attenuation and anisotropy

10.6.3 Temporal changes of coda


QC prior to earthquakes

Initially encouraging indications of a temporal change


in coda Qc before two large Chinese earthquakes were
reported by Jin and Aki, 1986. This was based on meas-
urements of coda decay rate for small local earthquakes
in the general area that is ‘preparing’ for a major earth-
quake. The authors referred to the significant difference
of duration of coda Qc from local earthquakes, which
may last several minutes in a stable area like Norway,
but which die out quickly in seismically active places
Figure 10.39 Stratigraphic section through the 1.4 km Tertiary like California. (Presumably, significant differences in
basin investigated by Clouser and Langston, 1991. attenuation, seismic Q and the rock quality Q-value
would also be registered in such contrasting regions).
During the three-year period preceding the Tangsham
earthquake in China (M  7.8, 1976), the coda Qc
was about 3 times lower than it was before or after this
period, i.e. attenuation was greater in the three years
before the event. Apparently a comparable change
occurred at the time of the Haicheng earthquake, also in
China. The authors emphasised that first appearances sug-
gested a change of focal depth. Tsujiura, 1978, reportedly
found that coda Qc for Japanese earthquakes were higher
when focal depths were at 100–160 km, than when at
40–80 km. Roecker et al., 1982, found coda Qc to be four
times higher when sampling 400 km depth than when
sampling less than 100 km depth.
The authors concluded that a coda Qc change by a fac-
tor of 2 to 3 would imply a change in focal depth of
about 100 km. They also discussed the possibility of
changes in predominant frequency affecting the coda Qc:

Qc  Qof m (10.10)
Figure 10.40 Qp-Qs relations determined for each station, from
spectral ratio slopes and travel times. Black circles
are laboratory and in situ data from other authors.
where 0  m  1. For the coda Qc to change by a fac-
Clouser and Langston, 1991. tor of 2 to 3 the frequency must change significantly
also. However, their analyses showed mean variations of
frequency of only 20%. They therefore concluded that
Vp the observed coda Qc change was due to a ‘change in
Qs   Qp (10.9) the property of the earth medium’, namely the opening
Vs of cracks or dilation. For the case of the referenced
Tangsham earthquake, the actual low coda Qc value
Using their average Vs/Vp and Vp/Vs ratios of 0.534 was 71 (in the period 1973 to 1976) and 200 (in the
and 1.87 respectively, the above equations gave the lower period 1969 to 1972). The P-wave velocity was report-
and upper limit curves shown in Figure 10.40, namely edly also anomalously low in the period 1973–1975 just
Qs  0.38 Qp and Qs  1.87 Qp. Their measurements preceding the earthquake. However, Jin and Aki, 1986,
using spectral ratios, and accounting for differences in preferred using coda Qc, since it covered an entire ellip-
basin thickness between each station, showed Qs vary- soidal region determined by the lapse time, whereas
ing from 10 to 70. monitoring of the Vp/Vs anomaly requires that locations
Seismic quality Q and attenuation at many scales 213

of source and receiver make the wave path go directly 10.6.4 Possible separation of
through the anomalous source region. attenuation into scattering
Peng et al., 1987, used several thousand seismograms and intrinsic mechanisms
of small earthquakes in the Mammouth Lakes area in
the USA to measure values of Qc from the decay of the Based on their own Mammouth Lakes data and on other
earthquake coda. They added a certain degree of sophis- case records, Peng et al., 1987 proposed that coda Qc
tication to the interpretation, by observing two opposite reduced within the aftershock zone but increased outside
trends at different distances from the source: this zone, when the main shock occurred. The observed
temporal changes to coda Qc suggested that hetero-
a) in the region near the main epicentre, measure- geneities were responsible for the scattering component.
ments of coda attenuation (Qc1) were higher for Presumably such features as lithological boundaries,
earthquakes that occurred after the main shock, clay-filled discontinuities and branch faults, could cause
than coda attenuation Qc1 of those that occurred scattering while the intrinsic component of attenuation
prior to the main shock (i.e. Qc was lower after the might be microcrack and joint related (i.e. thin cracks
main shock, due to greater attenuation). that were bearing high fluid pressures, that were there-
b) further away from the main epicentre, measure- fore sensitive to small stress changes. Peng et al., 1987
ments of coda attenuation Qc1 were lower after emphasised that separation of the intrinsic and scatter-
the main shock than before it (i.e. Qc was higher ing components of coda Qc introduced a severe non-
after the main shock, due to reduced attenuation. uniqueness in determining these parameters. However,
Could this perhaps be due to the reduced shear stress they suggest that probably a small number of strong scat-
at distance, and more ‘damage’ closer to the source? terers dominated the coda in a seismically active zone.
Such would also give lower rock qualities: quite logically Concerns about the relative magnitudes of intrinsic
closer to the source following the main shock. attenuation (Q1 intrinsic) and scattering (Q1 scat-
Peng et al., 1987, reviewed numerous cases of tempo- tering) has led to methods for separating these effects
ral changes of coda attenuation (Qc1) before and after by integrating the S-wave energy for successive time
large earthquakes. For brevity we will make a list without windows, as a function of earthquake hypocentral dis-
individual references, and refer the reader to the above tance. The method is described by Hoshiba et al., 1991.
authors. To aid interpretation with respect to coda quality Figure 10.23 shows how it was applied to Central
Qc as opposed to attenuation Qc1, we will list changes to California, Long Valley and Hawaiian earthquakes, giv-
coda Qc in the list that follows. The changes are therefore ing a separation of Qs1 scattering and Qs1 intrinsic,
inverted compared to the Peng et al., 1987 list: whose sum is equal to Qs1 total. Interestingly, when
adding two larger Qs intrinsic and Qs scattering compo-
1. 30% reduction in coda Qc before Hawaii earth- nents, the sum Qs total is less than either of the above,
quake, 1975 (M  7.2) since attenuation is being accumulated, not Qseis.
2. 20% reduction in coda Qc before 3 large Kuril- In general, for frequencies 6.0 Hz, scattering Qs1 was
Kamchatka earthquakes, (M  8.0) greater than intrinsic Qs1, whereas above 6.0 Hz the
3. 30% reduction in coda Qc (at 6 Hz) before Petatlan opposite applied. In all three regions, scattering Qs1 was
earthquake (M  7.6) strongly frequency dependent, decreasing proportionally
4. anomalous, low coda Qc before E. Vamanashi to frequency increase, or even faster. Intrinsic Qs1 was
earthquake, 1983 (M  6.0) considerably less frequency dependent. A concurrent
5. 300% reduction in coda Qc for 3 years before suggestion of a depth-dependent intrinsic Qs1, which
Tangsham earthquake, 1976 (M  7.8) increases with depth, apparently can reduce the dis-
6. 200% reduction in coda Qc for period before crepancy between theoretical predictions and observa-
Haicheng earthquake, 1975 (M  7.3) tions (Zeng et al., 1991).
Sato and Fehler, 1998, (Seismic wave propagation and
Following this list an increase in coda Qc is men- scattering in the Heterogeneous Earth, Springer Verlag &
tioned for a 2 to 3 year period before the Misasa earth- AIP Press), who collected regional seismic data from
quake in Japan, 1983 (M  6.2), and the authors also around the world, also showed the separation of the scat-
record the possibility that coda Qc changes without an tering and intrinsic attenuation. Clearly, and in almost all
obvious relation to a major earthquake. cases, the lower frequencies give the greatest attenuation
214 Rock quality, seismic velocity, attenuation and anisotropy

Figure 10.42 Regional seismic data assembled by Sato and Fehler,


1998, showing the relative magnitudes of the intrinsic
and scattering attenuation and its frequency depend-
ence, even between 1 and 10 Hz. KTJ  Kanto-
Tokai, Japan; KJ  Kanto, Japan; LV  Long Valley,
California; H  Hawaii; SC  Southern California;
SP  Southern Spain; and CC  Central California.

these total, scattering and intrinsic attenuations. (The so-


called ‘Albedo’ (Bo) is defined as the ratio Qsc1/Qtotal1).
Mayeda et al., 1991, had suggested that the more
complex regions of Hawaii and Long Valley required
models which incorporated heterogeneously distributed,
non-isotropic scatterers in a layered medium with
depth-dependent intrinsic Qs1. They suggested that
their results for scattering Qs1 indicated a length scale
of heterogeneity at least comparable to the wavelength
Figure 10.41 Plots showing total Qs1, together with scattering
for the lowest frequencies studied, of the order of a few
Qs1 and intrinsic Qs1, for earthquakes that
occurred between 1987 and 1990. Focal depths are
kilometres, presumably implying fault-size features, or
between 5 and 20 km and magnitudes are between perhaps lithological contacts.
1.5 and 3.6. Mayeda et al., 1992.

both from scattering and from intrinsic mechanisms. 10.6.5 Changed coda Q during
This seems to suggest the extreme importance of the seismic events
structural geology (joint sets, faulting), and of the con-
ducting properties of these larger scale, nearer the surface As we have seen, and perhaps confusingly for earthquake
features. Figure 10.42 shows the relative magnitudes of precursor analysis, some investigators have recorded
Seismic quality Q and attenuation at many scales 215

decreases in the seismic coda Qc value prior to major


earthquakes, while others have recorded decreases fol-
lowing the occurrence of a major earthquake.
Beroza et al., 1995, using nearly identical (doublet)
earthquakes in a pre-seismic, co-seismic and post-seis-
mic search for temporal changes in coda Qc, found a
nearly stable attenuation throughout the Loma Prieta,
California sequence. The main shock in October 1989
(M  6.9) nevertheless reportedly resulted in an almost
total stress drop, and effects on coda Q had clearly been
expected. In earlier studies, increases in coda Qc by as
much as 50% had been cited as precursors to large earth-
quakes. Unfortunately, changes in coda Qc by as much
as 50% had also been reported even in the absence of a
large seismic event.
To address possible effects of source size, geometric Figure 10.43 Temporal variation of coda Qc for a swarm of earth-
quakes with peak activity in August and September.
spreading and earthquake mechanisms and location,
Chen and Long, 2000.
most researchers estimate coda Qc from late coda, the part
of the seismic wave arriving after twice the S-wave arrival
time. The relatively long paths taken by waves in the late Chen and Long, 2000, found that their coda Qc val-
coda are assumed to sample a large volume of the crust ues were related to the locations of the earthquakes.
with a variety of takeoff angles from the source. Early events were mainly located in high coda Qc areas,
Using earthquake doublets, Beroza et al., 1995, while later earthquakes occurred preferentially in low
employed a technique for measuring coda Qc that was coda Qc regions, in other words there was a shifting of
assumed to be insensitive to geometric spreading, loca- hypocentres. (Rock quality or attenuation levels could
tion and source mechanism, because these factors are presumably change also in a given location, due to frac-
common to the two (doublet) events. Furthermore, turing events, but hypocenter migration was considered
both the early and late coda can be used; but the former the main effect). The coda attenuation Qc1 increased
is usually sensitive to the focal mechanism. from about 0.005 (Qc  200) to 0.010 (Qc  100)
The stability of coda Qc throughout the Loma Prieta from the early period to after the peak activity, due to
sequence was in sharp contrast to other studies (e.g. this assumed shift of hypocentres.
Peng et al., 1987), who reported larger pre-cursory A short distance (1 km) migration of hypocentres
changes in coda Qc for many (but not all) of the earth- generated this halving of magnitude of Qc, possibly
quakes that were analysed, as we saw earlier. related to migration from ‘geologically uniform’ gneiss to
Previously it had been assumed that the coda samples a an area with ‘mafic dikes and more complicated topo-
large volume. However, Chen and Long (2000) showed graphic relief ’. These dikes were interpreted as strong,
that temporal variations in the seismic coda Qc could in and inhomogeneously distributed scatterers. The authors
fact be explained by hypocentre migration over small dis- concluded that the normal assumption that coda scatter-
tances compared to the radius assumed to be sampled by ers (i.e. large discontinuities) are uniformly distributed
the coda. These authors used data from an earthquake may be one reason for earlier misinterpretations of tem-
swarm in Georgia, USA (Norris Lake Community). poral variations in coda Qc. An assumed homogeneity
They first noted the apparent temporal variation in of velocity structure would be another.
coda Qc during a five month period. Figure 10.43 The strong spatial variation of coda Qc (and by
shows that reductions occurred in the three orthogonal implication of the rock quality Qc-value also) does not
directions (vertical, NS, EW) during the period of max- eliminate the possibility that intrinsic attenuation and
imum seismic activity. Over 4000 shallow (1.2 km) scattering may also be affected by more subtle changes
earthquakes were recorded altogether. However for this in rock mass properties and the effects of stress change.
analysis, 108 earthquakes of similar size were chosen However, in rock engineering, we would certainly
that were uniformly distributed during the period of expect that spatial variations of rock mass quality Qc
the swarm activity. About 2 weeks before the peak of would tend to be greater than those caused by stress
seismic activity, the coda Qc began to reduce. change, especially in view of the moderate stress drops
216 Rock quality, seismic velocity, attenuation and anisotropy

that occur as a result of earthquakes (i.e. often just frac- January 1994. Despite magnitude (M 4.7 and M 4.6)
tions of, or a few MPa). events in 1992 and 1993, they found that Qc had not sys-
Following rock quality Qc values a little further, one tematically changed. Figure 10.44 however, shows various
can see from Figure 5.36 (Part I, Chapter 5), that higher interesting trends of the data, including a certain
rock quality Qc values imply higher P-wave velocities anisotropy regarding azimuth (graph c), and a clear dis-
and higher deformation moduli. Such areas would likely tance dependence (graph d) which may be related, as
be under highest stress, and have least permeability, so a observed earlier, with deeper sampling of the earth’s crust
spatial migration of hypocentres to lower seismic coda as distance increases, and therefore higher coda Qc values.
Qc areas, with lower rock mass quality Qc and lower Concerning the coda Qc magnitude and time period,
moduli and higher permeabilities, is entirely logical. Hellweg et al., 1995, suggest that there had been no sys-
Perhaps the ‘complicating’ factor of frequency depend- tematic change in the coda Qc. One may however observe
ence of seismic Q reviewed earlier, is another way of that if allowed to plot a least-squares (or other) best fit to
recognising a scale effect. Certainly the evaluation of the time/date data in graph (e), a certain reduction of
rock mass quality Qc could also be considered scale- Qc with time would be observed. However this would
dependent, since the inclusion of larger volumes of the apparently be invalid as the events cannot actually be
rock mass (including faults) will inevitably adversely compared with each other directly, as they are deter-
affect all the six Q parameters, resulting in lower over- mined from different length windows regarding each
all rock mass quality and the strong likelihood of seismogram. The authors proposed that Qc should always
greater attenuation. be measured from the same length window starting at
Hellweg et al., 1995, also used the Parkfield Dense the same lapse time regardless of the source location.
Seismograph Array (of the US Geological Survey) to esti- As a contrast to the coda Qc obtained from seismic-
mate coda Qc from up to 42 recordings for each earth- ally active areas, Kvamme and Havskov, 1989, deter-
quake that occurred. Coda Qc was determined in two mined the coda Qc in Southern Norway, finding values
frequency bands (4 to 8 Hz, and 8 to 16 Hz), from a tight at 10 Hz frequency to vary from 780 to 1530, for
cluster of 26 seismic events between December 1989 and source-to-station distances varying from 15 to 300 km.

Figure 10.44 Coda Qc from the 4–8 Hz band: a) coda Qc dependence on depth, b) coda Qc dependence of earthquake magnitude, c) coda
Qc dependence on azimuth, d) coda Qc dependence on epicentral distance (km assumed), and e) coda Qc as a function of
time. Dotted lines show the average of all coda Qc values, f ) coda Qc calculated for selected events with epicentres less than
30 km. Filled diamonds are measured from a 30 s window which starts at 2ts. Hollow diamonds are measured from a 30 s win-
dow which starts at a lapse time of 20 s. Hellweg et al., 1995.
Seismic quality Q and attenuation at many scales 217

A certain increase in coda Qc with window length was The authors used local earthquakes recorded from 1987
interpreted as increased Qc with depth, as in other stud- through 1996, and concentrated their analyses on high-
ies reviewed here. They considered the Norwegian quality data from the depth range 45 to 15 km. Values
measurements of coda Qc to be similar to values found of coda Qc1 were averaged over three recording sta-
in another shield area (Canada), but observed stronger tions for each earthquake. They divided the data into
frequency dependence as possible evidence of stronger two periods: 8 years before and 2 years after the major
scatterers in Southern Norway. Some of the paths shown (M  7.2) 1995 earthquake.
in Figure 10.45, certainly cross some major regional zones The average value of coda Qc1 increased after the
of weakness (i.e. the Oslo fjord), and regional faulting. major earthquake, especially for the lower frequency
They used window lengths of 5, 20, 30 and 40 s and bands between 1.5 and 4 Hz, as illustrated in Figure
observed variations of coda Qc with frequency: 10.47, and in Table 10.5. The authors emphasised that no
change in focal mechanism was reported, citing the fact
Qc  Qof m (10.11) that changes in epicentres, focal depths, or focal mecha-
nism can cause false temporal changes in coda Qc1. Even
and found (m) to be 1.15 for most of the Norwegian in a small 1  1  1 km volume, the value of Qc1
data. Comparison of their own and other frequency increased after the main shock at frequencies below 5 Hz,
dependent coda Qc are reproduced in Figure 10.46. suggesting that changed epicentres were not the cause of
Hiramatsu et al., 2000, reported temporal changes in the increased attenuation (and reduced seismic quality).
coda Qc1 in the Tamba region of Japan, to the north- The average depth remained in the 9 to 10 km range.
east of the main rupture zone of the 1995 M 7.2 The authors considered that a numerically estimated
Hyogoken Nanbu earthquake. This region has the change of shear stress at 10 km depth of only 0.02 MPa
densest distribution of Quarternary active faults in due to the Hyogo-ken Nanbu earthquake, was the
Japan, with very high seismic activity for several decades. cause of the increased attenuation. The sensitivity of
Qc1 to shear stress change was estimated to be 10
(MPa)1 at around 3 and 4 Hz frequency, which the

Figure 10.46 Coda Qc as a function of frequency for different


regions. A–Aleutian Islands, m  1.05, B–Carolina,
USA, m  0.94, C–New England, m  0.40,
Figure 10.45 Location of Southern Norway profiles for Kvamme D–Southern Norway, m  1.15, E–Canadian
and Havskov, 1989, spectral ratio analyses. The dot- shield, m  0.20, F–former Montenegro region,
ted line direction gave seismic Qp  575 from meas- Yugoslavia, m  1.00. (See Kvamme and Havskov,
urements by Kanestrøm and Haugland, 1971. 1989 for references).
218 Rock quality, seismic velocity, attenuation and anisotropy

Figure 10.47 Distinct increases in attenuation (coda Qc1) following the January 1995 Hyogo-ken Nanbu (M  7.2) earthquake.
Hiramatsu et al., 2000.

Table 10.5 Increase of average Qc1 and decrease of average Qc (n) for the activity periods I to V (shown in Figure 10.48).
following the 1995 Hyogo-ken Nanbu earthquake. All ten frequency bands (from 1.5 to 24 Hz) were
(Hiramatsu et al., 2000). analysed. As clearly shown in this figure, there was a clear
Frequency (Hz) Qc1 before Qc1 after Qc before Qc after reduction in the n-value in the years leading up to the
event, followed by a marked increase in the n-value, espe-
1.5 0.012 0.015 81.3 67.6 cially after the major event. As suggested by Kvamme and
2.0 0.011 0.013 91.2 77.6
Havskov, 1989, (and others), a greater density of scatter-
3.0 0.0076 0.0093 131.8 107.2
ers could be the logical cause of this greater sensitivity to
4.0 0.0054 0.0062 186.2 162.2
frequency, following the major earthquake.

authors suggested was much larger than the stress sensi- 10.6.6 Attenuation of damage due to
tivity of seismic velocity. However they also referred to acceleration
the fault or fracture dimensions of micro-earthquakes
(M  3) as about 400 m, consistent with the character- Mandal et al., 2001 interpreted 110 aftershocks following
istic length of scatterers of 300–600 m estimated from the Mw 6.4 Chamoli earthquake in the Garhwal
the wavelengths of 3 to 4 Hz which had the greatest Himalaya as propagating up-dip along a thrust plane from
influence on Qc1 increases. 20 to 2 km depth, the main shock having occurred at
They found in addition, that the frequency depend- 15 km depth. The region had five earthquakes exceeding
ence of coda Q1 varied with time. Using the propor- magnitude 6, and twelve exceeding magnitude 5 in just
tional to power of frequency expression: the 20th century, which was presumed to have caused a
high level of shallow crustal heterogeneity. They analysed
Q1 1 n
c  Qo f (10.12) 48 of the local earthquakes with magnitudes varying from
2.5 to 4.8, with recordings at nine digital stations, with
where Qo1 is the Qc1 value at 1 Hz and n is a constant, three-component seismometers, covering an area with a
they used the least squares method to determine values of 150 km radius. Sub-surface recording is not mentioned.
Seismic quality Q and attenuation at many scales 219

a zone V (potential M8 earthquake zone), despite the


intraplate location of this Kutch area, is evidenced by
the two 7.7 magnitude quakes of the last 200 years. The
authors suggested, as at Garhwal, that these were the
reasons for a high level of shallow crustal heterogeneity.
The coda Qc interpreted from 200 local earthquakes
of magnitudes varying from 3.0 to 4.6, in an area of
140 km radius, suggested Qo  102 f 0.98. On this
occasion, based on the attenuation curve estimated
for Qo  102, they estimated a ground acceleration at
240 km distance (from an assumed 1 g source) of 0.13 g,
almost agreeing with a 0.11 g accelerograph recording
at similar distance. To be read in conjunction with this
locally attenuating, seismically active crust, is the very
variable velocity-depth model for the uppermost 6 km
of the region given by Mandal et al.:
Figure 10.48 Frequency dependence of Qc1 increases over the 0.9 km 2.2 km/s
same 10-year period. Hiramatsu et al., 2000. The 0.2 km 4.8 km/s (basalt)
sharp rise after the event in period IV was referred to 2.6 km 3.5 km/s (Jurassic sediments)
in their figure.
0.1 km 4.8 km/s
2.2 km 5.9 km/s
The estimated Qo value (Qc at a frequency of 1 Hz)
at the various stations was on average only 30 at 1 to 10.6.7 Do microcracks or tectonic
3 Hz, indicating a strongly attenuating crust. Higher structure cause attenuation
Qc at higher frequencies were interpreted as propaga-
tion of back-scattered body waves through deeper parts When considering the mechanisms of attenuation of
of the lithosphere where less heterogeneities were the crust from a great earthquake, the decay of wave
expected. South of the main central thrust belt is a less amplitude with distance can clearly take many forms:
rigid, slightly metamorphosed sedimentary wedge, conversion of elastic energy into heat, scattering caused
while north of the thrust are more rigid highly meta- by heterogeneities such as changes of rock types, changes
morphosed crystalline rocks. The frequency depend- of velocities, sub-surface topography, and cracks and
ence of the Garhwal Himalaya coda Qc also gave a quite faults. The authors Mandal et al., 2004 also cite ‘sliding
‘steep’ curve in relation to Figure 10.46 data from other along grain boundaries’, but in the opinion of the
seismically active regions. Qc  30 f 1.2 was typical. writer this seems insignificant in relation to the, admit-
The author compared the Garhwal Himalaya Qo of tedly surface-magnified, ground deformation, with wide-
30, and its frequency component (n) of 1.21, with the spread uplift, deep cracks, liquefaction, ejection of water
Indian Peninsula Qo of 550 and n  0.84. From the and sand, even at 275 km distance, and collapse of
above high attenuation, the authors, in 2001, infer that 50 high-rise buildings in a city (Ahmedabad) 240 km
the acceleration (from a hypothesised 7.4 magnitude distant from the epicentre. There was damage even at
earthquake at Chamoli, with 1 g at the source), would 1,600 km distance on the SE coast of India – probably
decay to 50% at 20 km distance, and to 7% at 100 km, due to the high seismic Q of the intervening crustal plate.
suggesting less significant damage beyond 100 km if It seems intuitively impossible that micro-friction along
such should occur. microcracks, joints and faults, as an attenuation mech-
In a later paper, Mandal et al., 2004 report on the low anism does not exist in the crust. An earlier opinion
coda Qc interpreted from the devastating, intraplate Mw from the 1980s, based mostly on the basis of high fre-
7.7 earthquake that occurred in Gujarat in 2001, named quency, sub-micro-strain oscillation tests on intact rods
the Ghuj earthquake. On this occasion the distribution of sandstone, and continuum concepts of strain, deduced
of aftershocks defined a 45° dipping zone of some that friction was not a source of attenuation in the earth.
40  60 km in area. The presence of high seismicity, in Some geophysicists hold this viewpoint even today.
220 Rock quality, seismic velocity, attenuation and anisotropy

Due to the higher stiffness of grain boundaries and


microcracks – they have ‘high’ aspect ratios – it would
also seem that the ubiquitous rock joint, and its prob-
able several sets, could be a likely candidate for micro-
friction attenuation combined with squirt flow losses.
Attenuation caused by friction is certainly easy to envis-
age in the case of an earthquake induced, 8.5 m rupture
at 23 km depth, reducing to about 1 m near the surface,
where there was uplift of the southern side of a major
fault. Neighbouring rock masses resisting such motion
are clearly absorbing frictional energy, suffering micro- (a)
displacement discontinuities in the process.
If joint and fault compliances that fit attenuation
data somehow ‘acquire’ realistic magnitudes and units
of Pa1 m (or the more familiar MPa/mm of stiffness)
during even minor man-made seismic or sonic explo-
ration, how can friction-caused attenuation, of however
small magnitude, actually be avoided in the case of nat-
ural seismic events? Such ‘displacement-discontinuity’
events in a jointed rock mass are surely a part of the
intrinsic attenuation. That scattering occurs from these
features as well, seems hardly justification for excluding
compliance effects from intrinsic mechanisms, as
implied in some interesting analyses of fault-related
attenuation. (b)
From another seismically active region, the authors
Figure 10.49 Qp interpreted from seismic events, in terms of
Akinci et al., 2004, present highly attenuated data from
a) distance and b) magnitude, in the western and
the North Anatolian Fault Zone in Turkey, and com- eastern portions of the North Anatolian Fault Zone.
pare this with a sediment influenced region of Southern The NAFZ stretches 1400 km from eastern Turkey
Germany. The NAFZ is an intra-continental transform to the Aegean. Lack of sediments at all the receiver
fault boundary between the Eurasian Plate in the north, stations suggested to the authors that the low Qp
and the Anatolian block in the south. This major strike- applied to crystalline basement rocks of Triassic
slip fault extends through many segments, for about age. Akinci et al., 2004. Sub-site attenuation also
1400 km, from a triple plate junction in eastern Turkey seems a possible source of the lowest values.
to the Aegean in the west. Significant levels of jointing/fracturing/faulting must
Figure 10.49 and 10.50 show the results of the authors be assumed in view of the low values of Qp.
P-wave Qp analyses for the western and eastern portions
of the NAFZ, with Qp versus distance data, and Qp ver- range of expected rock mass deformation moduli for
sus seismic event magnitude. Since there is little sedi- the following conditions:
ment beneath the stations in either region, the low Qp
results possibly apply to crystalline basement, a lot of it 1. Near-surface (e.g. immediately below station): rock
of Triassic age, although ‘site effects’ from a more atten- with a compression strength more than 10 MPa
uating near-surface rock mass beneath the recording and a rock mass Q-value in excess of 1 (heavily
stations, should perhaps not be neglected here. Most of jointed), giving rock mass Qc  0.1. Figure 10.54,
the seismic Q results lie between about 4 and 40, sug- and 13.60 in Chapter 13 show Emass  5 GPa (as
gesting, on the basis of an intuitive Qseis – Qrock quality per minimum seismic Qp in Figure 10.49a).
relation, that fracturing/jointing and faulting must be 2. At 1 km depth, rock with a compressive strength
extensive in this tectonically disturbed region. of 100 to 300 MPa, and a rock mass Q-value of
The range of data for Qp shown in Figure 10.49, for- 100 (jointed but massive) to 1000 (almost without
tuitously or for scientific reasons, exactly matches the joints – or completely closed joints) giving Qc from
Seismic quality Q and attenuation at many scales 221

depth than ever directly measured. Linked Qp-Emass val-


ues in a range of 100–400 (also GPa) would appear rea-
sonable, but this is pure speculation of course.

10.6.8 Down-the-well seismometers


to minimise site effects

(a) Knowledge of attenuation magnitudes in the upper few


kilometres of the earth’s crust is clearly an essential
ingredient in modern seismic hazard analysis. Earlier
studies using only surface seismometers were limited by
relatively high noise levels and by the strong attenuation
at shallow depth, preventing high frequency signals
from being recorded.
The installation of borehole seismometers in more
recent years has greatly improved knowledge of the near-
surface attenuation of some individual sites. Abercrombie,
1998 pointed out that seismograms recorded at the
earth’s surface are contaminated by both seismic and
(b)
man-made noise. The frequency range of observed sig-
nals tends to be limited to below a few tens of Hertz at
Figure 10.50 a) Separation of sediment and basement Qp values by most surface sites. This makes it difficult to link obser-
proportioning of the respective velocities and dis- vations of attenuation in the real earth with the much
tances, following Hough et al., 1988. b) Effect of cor- smaller scale and much higher frequency laboratory
rection on the corrected Qp values for the Southern studies. The ideal situation for recovering uncontami-
German seismic data. Akinci et al., 2004.
nated measurements is installation of wide-bandwidth
seismograms down deep boreholes, where there is low
100 to 3000. Figure 10.54, and 13.60 in Chapter 13 background noise. The problem of strong attenuation
show Emass from approx. 50 to 140 GPa (according to in the near-surface is thereby avoided.
empirical equations: specifically 46 to 144 GPa). As part of the Californian earthquake prediction pro-
3. Note the great body of seismic Qp data between 10 gramme, a region of the San Andreas fault near Parkfield,
and 40, suggesting domination of near-surface, sub- California was equipped with an array of seismic instru-
station, heavily jointed and sometimes faulted rock mentation in boreholes averaging about 250 m depth,
mass. Refer to ‘minor fault’ curve in Figure 10.54, with a more extensive downhole array both at 1.5 km
at depth from 50 to 1000 m: Qp data is matched and almost 3 km in the Cajon Pass deep well, some
by moduli of 10 to 40 GPa. 1.5 km NE of the fault. Deep instruments reportedly
failed soon after installation in the 5150 ft. deep Varian
By way of comparison of a seismically active region well (Malin et al., 1987). In this area, quite different
with one that was less so, the authors also presented their basement rocks are juxtaposed by the fault, with higher
Qp results for a region of southern Germany. As indi- velocities to the SW than to the NE.
cated in Figure 10.50, they made a correction for the low Deployment of closely spaced seismometers has report-
Q sedimentary cover of some 2 km thickness, which had edly shown extensive variation in recorded amplitude, fre-
a low range of Qp from 6 to 10, thereby revealing the cor- quency content, and coda duration over short distances,
rected Qp for the 8 to 10 km of basement of between 100 meaning that the ‘site effect’ (the near-surface, near-
and 500. The fact that the uncorrected 2 km deep data receiver rock), can have a strong effect on earthquake
match the above, while the 8 to 10 km of basement recordings. Amplification in the low velocity, low density
exceeds the above Qp – Emass (potential) link, can per- near surface, scattering, and resonance within shallow lay-
haps be ascribed to the need for an extreme extrapolation ers, plus attenuation of high frequency energy, also play a
of deformation moduli to an order of magnitude greater role in earthquake damage.
222 Rock quality, seismic velocity, attenuation and anisotropy

Seismic Q (i.e. Qp unless Qs is specified) is now known


to be as low as 10 in the upper 100 m of many sites, in
varying rock types. Attenuation at shallow depths in fact
appears to exhibit little dependence on rock type. This,
to a degree, is also basically true for the rock mass quality
Q. According to reviews by Abercrombie, 1998, and
Abercrombie (2000), the instrumented deeper boreholes
show seismic Q increasing with depth to about 100
between 1 and 2 km, and reaching about 1000 at greater
depths. Adams and Abercrombie, 1998, found seismic
Q  1000 below 2.5 km depth at Cajon Pass. Studies
using both direct and coda waves down-hole-recorded
at over 2 km depth, have shown seismic Q to be high
(1000) at seismogenic depths in California. Hough
et al., 1998, found seismic Q  1000 at 5 km depth at
Anza, in California.
A useful comparison of the relative magnitudes of
attenuation in the Varion and Cajon Pass wells given by
Abercrombie, 1998, is reproduced in Figure 10.51.
Despite the difference in rock types, the seismic Q pro-
files are essentially similar. However at greater depths, Figure 10.51 A comparison of attenuation-depth profiles for Qp
the thicker and lower velocity Franciscan rocks to the and Qs in the Varian well, Parkfield, and in the
NE of the San Andreas fault at Parkfield, appear to give Cajon Pass well, S. California. Both sites show very
lower Qp and Qs than the higher velocity Mesozoic low Qseis near the surface, increasing with depth. As
crystalline basement rocks at Cajon Pass. Abercrombie’s observed elsewhere in this chapter, similarity to
analyses of the spectral ratios at different recording depth-modulus trends for jointed rock masses is
striking. Abercrombie, 1998.
depths, having linear slopes on a log-linear plot, sug-
gested that the near-surface seismic Q was almost inde- Table 10.6 A velocity and attenuation model for either side of the
pendent of frequency, in the frequency range of about San Andreas Fault (SAF), with average seismic Q
2 to 100 Hz. inversion results for the upper 5 km, assuming a Qp
A velocity model crossing the fault at right angles is value of 1000 below 5 km (Ambercrombie, 2000).
shown in Table 10.6, based on Abercrombie, 2000. Qp
Depth Vp (SW) SAF Vp (NE)
and Qs estimates down the Varion well on the NE side
of the fault are given in Table 10.7. 0–2 km 4.7 SAF 4.0
Earthquake sources reportedly lie within the steep 2–5 km 5.9 SAF 5.15
velocity gradient separating the higher velocity SW side 5–10 km 6.1 SAF 5.35
10 km 6.5 SAF 6.0
of the fault from the lower velocity NE side. This is con-
sistent with micro-seismic and rock burst observations Qp (SW) Qp (NE)
reported in mining and deep tunnelling projects referred 0–5 km 199 49 99
to in Chapter 7. Greatest AE activity occurs where vel- 5–12 km 1000 – 1000
ocity gradients are highest, as in such parts of deep gold
mines, according to interpretation of seismic AE arrays. Qs (SW) Qs (NE)
Abercrombie, 1997 calculated that an earthquake –5 km 236 78 84
recorded at the Cajon Pass wellhead, with a hypocentral 5–12 km 1000 – 1000
distance of 15 km, would suffer 90% attenuation in the
[SAF  San Andreas Fault: Qp  50, Qs  80 (top 5 km)].
upper 3 km, 80% in the upper 1.5 km, and 50% in the
upper 300 m, at frequencies of a few Hertz and above. would occur in the upper 1 km: the relatively lower and
Equivalent calculations using the data in Figure 10.51, deeper seismic Q magnitudes at this well constituting a
for the same earthquake hypocentral distance at the Varian form of ‘site effect’, with most attenuation occurring at
well, suggested that less than 50% of the attenuation greater depth than 1 km.
Seismic quality Q and attenuation at many scales 223

Table 10.7 Qp and Qs estimates at various depth intervals in the


Varian (VAR) well from Ambercrombie (2000) and
Jongmans and Malin, 1995.

Depth interval (m) Qs Qp

0–298 7–10 20
298–938 31 30
572–938 53–94 55
0–938 18, 33–45 33

Analysis of seven earthquakes recorded at 2.5 km


depth in the Cajon Pass well, reported by Learey and
Abercrombie, 1994, found a weak increase in Qs with
frequency (500 at 10 Hz, increasing to 1200 at Figure 10.52 A compilation of Qs estimates and their frequency
100 Hz). They found that intrinsic attenuation was the dependence between 1 and 10 Hz. Abercrombie,
dominant mechanism, because of the ratio Qscatter  1998, from Adams and Abercrombie, 1998, Leary
10 Qintrinsic, at all observed frequencies. and Abercrombie, 1994, and Kinoshita, 1994.
(Relevant here, perhaps, is a contribution to the ques-
tion of frequency-dependence by Van Der Baan, 2002. He
used wave localization theory to show that ‘constant’ Q Abercrombie, 1998 cited several studies from other
might be due to apparent attenuation due to scattering tectonic areas in the Western USA and from Japan,
losses, if the earth displayed fractal characteristics over a which showed frequency dependence of Qs between
certain range of scales, thereby creating an absorption frequencies of 1 to 10 Hz, with a levelling off of fre-
band. The author drew attention to the usual constant quency dependence above 5 or 10 Hz. The lesser fre-
Q at low frequencies, and some form of positive power law quency dependence above 5 to 10 Hz (see compilation in
dependence at higher frequencies. Constant Q for exam- Figure 10.52), is suggested as a possible reflection of
ple between periods of 1 hour and 10 seconds (0.3  104 changes in the nature of the crust at scale lengths of a few
to 0.1 Hz), is followed by an increase for f 1 Hz. The hundred metres (an REV effect?), or that it could be due
author suggested that it may be very difficult to distin- to an artefact of the models, which at the outset, assume
guish between intrinsic and scattering attenuation body or surface waves that are isotropically attenuated
using only frequency-dependent Q measurements. He (no lateral or depth variation) – reasonable with at-depth
also claimed that, in most cases, the common assump- recordings, but not with many surface recordings.
tion that the scattering and intrinsic effects could be The level of current tectonic activity, a thinner warmer
separated, by allowing only for frequency-dependent crust, the presence of large crustal faults characterized
scattering, could be invalid. His numerical modelling by low velocity zones, were each referred to, as possible
indicated that in the case of non-perfect fractals, a con- reasons for the marked frequency-dependence between
stant Q only occurred within a given frequency band, 1 and 10 Hz. There has been much made of the strong
and that Q became proportional to f  for higher fre- and continued frequency dependence indicated by earl-
quencies, as in fact observed in most cases.) ier recordings of Qp and Qs from seismic coda, but
Adams and Abercrombie, 1998, analysed more than most likely such records were not recorded at various
100 earthquakes recorded at a range of depths in the Cajon depths below the surface, and therefore had numerous
Pass borehole. They employed a multiple time lapse potential ‘site effects’.
method to determine the relative contributions of intrinsic Abercrombie, 2000, compared recordings of nine local
and scattering attenuation concerning Qs and its fre- earthquakes at seven different depths down the Varian A1
quency dependence. They confirmed only a weak fre- well (0, 24, 298, 572, 877, 907 and 938 m) and at the
quency dependence (Qs  800 at 10 Hz, increasing to other borehole network in the Parkfield area (the high res-
1500 at 100 Hz, similar to the above). These authors olution seismic network – HRSN). The layout of these
also found intrinsic attenuation to be the dominant mech- measurement locations in relation to the various branches
anism, compared with scattering, at seismogenic depths. of the San Andreas fault is shown in Figure 10.53. In the
224 Rock quality, seismic velocity, attenuation and anisotropy

listed from Abercrombie, 1998 discussion of this topic,


as follows:

1. Fracture densities at outcrops and in cores suggest


perhaps an order of magnitude decrease at 500 m
or more depth (this of course is variable).
2. The presence of fractures (or joints) in the upper
kilometres, and the moderate pressures, suggest
that friction may be a dominant mechanism of
intrinsic attenuation.
3. Analysis of seven earthquakes recorded at 2.5 km
depth in the Cajon Pass well, had reported
Qscatter  10 Qintrinsic, at all observed frequencies.
4. Joints and fractures are also major scatterers of seis-
mic energy, and reduction of scattering losses with
depth would be expected due to their reduced fre-
quency and greater closure with higher stress.
Figure 10.53 Parkfield, California high resolution seismic network
(triangles) and Varian array (VAR). Earthquake epi-
5. Mining induced seismic events (Spottiswood, 1993)
centres used in the Abercrombie, 2000 analysis of at 2 to 3 km in South Africa show Qseis of about
depth dependent and spatially variable attenuation, 1000 through ‘solid rock’ (i.e. probably more mas-
across the San Andreas fault are shown as circles. The sive quartzites), while close to stopes through highly
station depths follow the station initials. fractured ground (see the velocity EDZ of Chapter 7)
could be as low as 20.
analyses performed, Abercrombie, 2000, assumed that
The above mechanisms of attenuation, and the
attenuation was exponential and frequency-independent
depth-dependent seismic Q data, continue to support
The San Andreas fault zone was confirmed as being
the idea that seismic Q gives a strong reflection of rock
a strongly attenuating zone with Qp averaging 50 over
mass characteristics, with low values of seismic Q
the depth range 0 to 5 km. Seismic Q appears to increase
corresponding to the poorer, more jointed, more open
most rapidly at shallow depths, as is also the case for Vp,
structure that is typical of shallower rock, and perhaps
for the rock quality Q or Qc, and therefore for the rock
present beneath recording stations that are supposedly
mass deformation modulus. The attenuation to the
without ‘site effects’, i.e. those founded at the surface,
deepest instrument (0.9 km) on the lower velocity NE
on crystalline basement rocks.
side of the fault was comparable to the attenuation to
It seems increasingly reasonable to assume that the
the instrument at only 200 m depth on the higher
typical features of rock masses that make them variable
velocity SW side of the fault. Abercrombie emphasised
media for engineering construction (foundations for
the value of multiple-depth monitoring and lateral
large buildings and dams, tunnels and rock caverns),
arrays, in improving the level of characterization.
make them also ‘variable media’ and variably attenuat-
If Qp or Qs were linked to rock quality terms, such as
ing, in terms of seismic Q. Rock mass quality is import-
the rock mass deformation modulus, even better under-
ant, perhaps at all depths where jointing is present or
standing might perhaps be achieved of the physical
slightly open, because of implicit deformation moduli
nature of the rock masses close to and within the fault.
links to Qp variation with depth in the upper 2–3 km.
The range of Qp from the top 300 m to 2 km depth,
Up to this ‘mining depth’, empirical rock mass data is
seen in Figure 10.51 (10 to 180) is a near match to
still acquirable, or already existing.
entirely feasible deformation moduli (expressed in GPa),
One may speculate about what the rock mass Q-
as we saw for the Turkish data.
value distribution would be at the various depths in the
Varian well, if Q-logging had been performed on recov-
10.6.9 Rock mass quality parallels ered core, as done regularly in (usually shallower)
nuclear waste related projects, such as the 9 km of Q-
The potential reasons behind the strong attenuation in histogram core logging, sometimes to 1.6 km depth,
near-surface rock masses are numerous, and can be performed at UK Nirex’s Sellafield site (Barton et al.,
Seismic quality Q and attenuation at many scales 225

Table 10.8 Three progressively worsening rock mass qualities and their predicted (near-surface, hypothetical recording station) properties.
Consult tables in Appendix A for explanation of the selected ratings.

RQD Jn Jr Ja Jw SRF Q c (MPa) Qc Vp (km/s) M GPa K (m/s)

30 15 1 4 1 5 0.1 50 0.05 2.2 3.7 2  106


60 12 1.5 2 0.66 2.5 1 100 1 3.5 10 107
90 6 2 1 0.66 1 20 150 30 5.0 31 3.3  109

Note: Near-surface: Vp  3.5  log Q km/s, M(Emass)  10 Q1/3 GPa, K  1/Qc  107 km/s (Barton, 2002). A significant degree of
anisotropy can be provided if desirable or relevant, by using oriented RQDo and values of Jr and Ja perpendicular to the loading or (dynamic)
testing direction. The effects of anisotropic stresses or the effects of increased depth, and the effects of matrix porosity on Vp and M can be
handled using the equivalent depth and porosity corrections in Figure 10.54.

1992), and the Q-histogram logging performed at four Further insight into the meaning of these Q-parame-
1000 m deep SKB wells in Sweden at Forsmark and ters and their link to engineering parameters may be
Simpevarp in 2003. given by recording the fact that the first two terms
One cannot help but wonder whether the low seis- RQD/Jn, describing joint structure (the potential scat-
mic Q, also recorded elsewhere at crystalline sites, such tering component), has a maximum range from 100/
as the hundreds of Qp values lying between 4 and 40, 0.5  200 (massive rock without joints), through
recorded at the NAFZ (Figure 10.49a) from 5 to 30 km 45/9  5 (closely spaced joints in three sets), to 10/20 
distant events, are also a typical reflection of a strongly 0.5 (‘earth-like’, crushed rock). This pair of parameters
jointed, tectonically disturbed rock mass, also present alone, has by chance, a certain ‘familiarity’ in relation to
beneath the recording sites, and sampled by the lower relevant seismic Q magnitudes reviewed earlier.
frequency wave lengths. Sophistication, related to shear strength, is added by the
It is very easy to imagine a potential range of next two Q-value terms Jr/Ja, describing, with some level
rock quality Q, composed of the following typical Q- of accuracy, the friction coefficient (a potential intrinsic
parameter ratings, in such site locations (see Appendix attenuation component). (see Appendix A for rating
A for description of the ratings used to quantify the six descriptions, and for a graphic presentation of Jr/Ja 
).
parameters): Table 10.8 shows some hypothetical constructions of
typical near-surface Q-parameter ratings for potential
recording-site qualities. Such Q-parameter ratings have
RQD J J been applied on thousands of engineering sites world-
Q   r  w
Jn Ja SRF wide, and the rock engineering profession is as familiar
45  90 0.5  2 0.66  1 (10.13) with ‘Q  1’ as the geophysics profession is familiar
  
9  15 24 1  2.5 with ‘Qp  10’. Interestingly, and frustratingly, both
numbers can remain constant when actually composed
of different contributions from the various components.
(i.e. three or four sets of joints, spacing typically In rock engineering it is therefore good practice to
15–50 cm, one or more sets possibly slickensided or quote all six assumed parameter values. In geophysics it
smooth-undulating, with weathered or clay-smeared would be excellent if at least the near-surface scattering
joint walls, dry or partly saturated with water, with typ- (caused by RQD/Jn) and intrinsic components (caused by
ical shallow (25 m, or 100 m) near-surface, low- Jr/Ja and Jw) could be separated, as increasingly seen with
stress characteristics.) more deeply-acquired data. The hidden ubiquitous
From the above example, we obtain Q  0.1  10 microcracks have a ‘multiple role’ in increasing both the
(quality described as ‘very poor’ – ‘poor’ – ‘fair’). This attenuation components to varying degrees, but this is
calculated range shows, obviously, that Q rock quality  possibly masked by significant near-surface jointing, since
Q seismic, but when Q rock quality is used to estimate the of potentially stronger effect. Microcracks also have a role
rock mass deformation modulus Emass, values of from in increasing the effect of weathering, thereby reducing
about 2 to 30 GPa would be obtained with an appro- RQD and Jw, and increasing Ja, and possibly SRF – due
priate 10 to 100 MPa range of UCS, and the specified to loosening, all of which , it is suggested, cause a reduc-
25 to 100 m depth. tion in both of these Q values (Qseis and Qrock).
226 Rock quality, seismic velocity, attenuation and anisotropy

Figure 10.54 Rock engineering parallels to seismic Q. (See Chapters 5 and 9). The rock mass quality Q, with normalization to Qc to
account for weak or strong rock, appears to follow the trends of seismic Q values (with rock quality Qc  Qseis), Deformation
modulus M (Emass) expressed in GPa appears to match seismic Qp quite closely. Strongly attenuating fault zones, almost
‘invisible’ to seismic velocity, are seen to ‘maintain’ an apparently attenuating level of rock quality Qc at 1 km depth. It has been
observed that shallow Californian earthquakes are never found in regions with Vp  6.3 km/s. With appropriate stress
correction, this implies some degree of jointing in the neighbourhood of causative faults.

While microcracks are vitally important for attenu- 10.7 Attenuation across continents
ation in laboratory samples, they should be less so in the
near-surface rock mass, where jointing and weathering- As introduction to this section, concerning attenuation
induced porosity may readily dominate attenuation. across continents, it is appropriate to reproduce a well-
Some microcracks have also become macro-cracks in known diagram of plate tectonics. That reproduced in
this zone. Joint sets in petroleum reservoirs at 2 to 5 Figure 10.55 is from Isacks et al., 1968, as presented in
kilometres depth remain a major source of attenuation Kearey and Vine, 1996. It shows the classic subduction
(also polarizing shear waves), but microcracks presum- of the lithosphere into the asthenosphere at opposite
ably still contribute to the local attenuation caused by edges of plates.
the fluids. Any remaining pressure-resistant pore space
and higher aspect-ratio microcracks at 5 to 10 km
depth, may contribute to the strongly declined attenua- 10.7.1 Plate tectonics, sub-duction
tion, but there will still be scattering from major faults zones and seismic Q
and eventual rock boundaries. Jointing, usually assumed
to have ceased to exist at great depth, may in fact be The two-dimensional cross-section of the main features
present in the form of minor faulting, thereby explain- of an oceanic subduction zone, shown in Figure 10.56,
ing the maintained (but low) permeability. See the last is reproduced from Kearey and Vine, 1996. The
sections of Chapter 16 for discussion of these aspects. Benioff zone shown in this figure is the source of deep
Seismic quality Q and attenuation at many scales 227

Figure 10.55 Block diagram summarising the main features of plate


tectonics motions. The arrows shown in the astheno-
Figure 10.56 Cross-section showing main features of an oceanic
sphere represent possible complimentary flow in the
subduction zone. From Kearey and Vine, 1996.
mantle, in relation to the lithosphere. From Isacks
et al., 1968, reproduced in Kearey and Vine, 1996.

earthquakes, with the maximum stress direction follow-


ing the dip of this zone, as indicated in more detail in
Figure 10.57.
The ‘extremely low Q’ region shown in the figure
(Qp of 50 or less) lying in the mantle above the deep
earthquake zone appears to be a common feature above
such seismic activity. From a rock mechanics point of
view, it is reasonable to suppose that this is a zone of
high shear and tensile stress, with tectonic disturbance
of the strata as a result.
Within the original theory of plate tectonics, plates Figure 10.57 Schematic section crossing the Tonga trench show-
were considered to be internally rigid and to act as ing the high and low seismic attenuation regions,
extremely efficient stress guides. Supposedly a stress and the high seismic Q tongue of sinking (and
applied to one margin of a plate was transmitted to its thrusting) lithosphere, which is the source of deep
opposite margin with no deformation of the plate’s earthquakes. Bott, 1982. The continued application
interior (Kearey and Vine, 1996). The plates may be 80 of seismic Q, to hundreds of kilometres depth, is
to 150 km thick and thousands of kilometres in width. evident, and there is clearly structural/tectonic logic
In fact there are obviously some locations where in the location of many of the adjectives: extremely
intraplate deformation does occur, such as the thrust low, high etc.
faulting in intraplate mountain belts. In view of the fric-
tion to be overcome, from an engineering viewpoint it zones is known from analysis of the different seismic
would seem clear that the thrust at one side of the plate arrivals. High seismic Q travel paths suffer little attenu-
must be more than at the opposite margin, and the ation and represent stronger (and stiffer) rock. In the
variable modulus of deformation, albeit high, would case of the Tonga trench which has ocean depths of
then seem to have relevance in the overall behaviour. some 8 km, seismic waves passing up the length of the
It has long been suspected, e.g. Molnar and Oliver, Benioff zone appear to pass through a region of high
1969, that high values of Qseis correspond to regions of seismic Q (about 1000) while those travelling to lateral
high strength and high velocity, while low values corres- recorders pass through a more normal region of low
pond to low strength and low velocity, possibly associated seismic Q (about 150).
with high temperature and adjacent volcanic activity. A The zone of very high attenuation (or extremely low
long time ago, Daly, 1940, had correctly defined crustal seismic Q of about 50) identified in the uppermost
strength as ‘enduring resistance to shear stress with a mantle above the down-going slab, is a region about
limiting value’. 300 km wide (Barazangi and Isacks, 1971). It is implied
The Benioff zone, which is the source of numerous that this is much weaker (and less stiff ) than elsewhere.
earthquakes even down to a depth of as much as In the Basin and Range province of the Western USA
700 km, often dips at about 45°. The structure of such where there is also anomalously low Qseis in the upper
228 Rock quality, seismic velocity, attenuation and anisotropy

mantle, extensive normal faulting has occurred in a ten-


sional stress regime.
In the foregoing summary of estimates of Qseis in the
region of subduction zones, we have seen values of Qseis
ranging from high values in deep Benioff zones
(1000), low value in more normal (shallower) regions
(150), and extremely low values in the uppermost
mantle above the down-going slab (50); presumably
in the region of the pull-apart-basin.

10.7.2 Young and old oceanic


lithosphere

Noting that the coda Qc decay rate was supposedly


Figure 10.58 Coda Qc1 from coda analyses cited by Jin et al.,
independent of the source-receiver path, yet reduced by
1985, spanning several sites in the USA, Alaska,
an order of magnitude from stable continental upper Japan, Guam, China, Afghanistan, Iceland and Italy.
crust to active tectonic areas, Jin et al., 1985, investi- See Jin et al., 1985 for individual references. Note
gated coda Qc from 22 local earthquakes recorded on low-frequency, low seismic Qc trend, from suspected
islands located in young and old oceanic lithosphere. shallow crust sampling.
The authors found that coda Qc values increased
with frequency proportional to f n, where n ranged data and n was 0.61 for the younger AKU Iceland data,
from 0.46 to 0.61, the higher values being for younger as indicated graphically in Figure 10.58
oceanic lithosphere. This was similar to continental Their conclusions from a careful analysis of the data
regions where the active, tectonic, western side of the trends were that:
USA also show higher n values. Crustal coda Qc values
1. At frequencies from about 0.5–1 Hz, the average coda
at a frequency of 1 Hz reportedly increase from about
Qc values increased from the youngest oceanic lith-
140–200 in the western US to around 1300 in the
osphere to the oldest by a factor or 2.
stable central US, as we shall review shortly.
2. The regional change on the continents follows the
Figure 10.58 shows a large collection of data for coda
same trend, but the difference could be as much as
Q1c (attenuation) versus frequency of measurement,
a factor of 10.
which includes the island data of Jin et al., 1985
3. The coda Qc values from the old oceanic litho-
(termed ‘this study’) and numerous other authors’
sphere were therefore low in comparison to those
data assembled by R.S. Wu). See Jin et al., 1985 for a
obtained from stable parts of the continents.
complete listing of the sources of data. Most of the geo-
graphic areas are marked on the figure. The AKU and
GUA data are from Akureyri (Iceland) and the Mariana 10.7.3 Lateral and depth variation of
Islands of Guam. seismic Q and seismic velocity
The coda Qc data shown in Figure 10.58 demonstrate
largest differences at low frequencies, while all the trends Solomon and Toksöz, 1970, were among the early
converge at higher frequencies. Coda Qc tended to researchers who noted the significant lateral variations
increase beyond 1000 when frequency was greater than in attenuation. They referred to Oliver and Isacks,
about 20 Hz. Jin et al., 1985 proposed the equation: 1967 experiences of Qseis changing by an order of
magnitude over lateral distances of a few 10s of kilo-
( )
n metres. Demonstrable regional differences in crust and
Q  Qo f
fo (10.14) upper mantle Qseis are also found in the continental
USA, for example 200 in the Western seismic belt and
where fo is 1 Hz, and Qo is coda Qc at 1 Hz. They found 1000 in the east-central US, as will be reviewed shortly.
a good linear relation between log Q (i.e. coda Qc) and Variation of Vp and Vs with depth and lateral location
log f. Gradient n was 0.346 for the older GUA (Guam) in the crust, and corresponding Qp and Qs increases
Seismic quality Q and attenuation at many scales 229

Figure 10.59 Deep seismic refraction results showing depth variation of a) Vp and seismic Qp, b) Vs (and Vp/Vs) and seismic Qs. Iwasaki
et al., 1994.

with depth, are shown in Figure 10.59a and b, from from 3.1 to 5.4 km/s. The velocity structure below this
Iwasaki et al., 1994, from an extensive seismic refrac- showed lateral variation (as must surely be expected
tion experiment conducted on a 194 km N-S line over a 190 km profile), with successive increases in Vp
across the Kitakami massif of E. Northern Honshu in and seismic Qp as depth increased to 35 km. The Moho
Japan. A Jurassic accretionary complex lies to the north, occurs at about 32 to 34 km depth at the base of the
and pre-Silurian and Silurian-lower Cretaceous marine selected profile.
sediments lie to the south. Inspection of the variation of Vp and seismic Qp as
In these studies, the uppermost crust was covered depth increases was typically as in the simplified table
with a thin (0.5 to 1.0 km) ‘surface layer’ with Vp ranging shown on the next page.
230 Rock quality, seismic velocity, attenuation and anisotropy

Note the possibility of tentative extrapolation to 2 10.7.4 Cross-continent Lg coda


(and 5) km depth at the top of Figure 10.54. The impli- Q variations and their
cation of the data from Iwasaki et al., 1994, is of a sig- explanation
nificantly jointed upper 1 km, with low rock quality
Q-values. Beyond 2 km depth we move outside the area We will conclude this section on seismic quality and
of empirically derivable deformation moduli, so the attenuation, with continental broad-scale coda Qc
seismic Qp-Emass (GPa) similarity breaks down because variation first across Eurasia, then South America and
of lack of empirical data for Emass. the United States. Mitchell et al., 1997 showed that at
Qp beneath two of the world’s four major continen- a frequency of 1 Hz, the coda Q (from the so-called Lg
tal rifts (the Rio Grande, SW USA, and the East coda) varied between 200 and about 1000, with lowest
African rift) were estimated from 1000 km and 600 km values in the orogenic belt formed by the collision of
linear arrays during the 1980s. With Qp for the crust African/Arabian and Indian plates with the Eurasian
taken as 480 to 650, Qp values for the Rio Grande rift plate (Figure 10.60). Low values are also found in the
were 95 and 100 respectively. The value of Qp for the Arabian Peninsula (Qc  350–500), which is a region
more strongly upwarped asthenosphere beneath the of recent uplift, extension and volcanism.
East African rift was only 27 by comparison. In general High coda Qc (800) were found, as expected from
terms the asthenosphere is believed to have lower vel- earlier results, beneath three shield areas (East European,
ocity, higher attenuation and the possible presence of Siberian and Indian) and beneath the oldest portion of
partial melt, which significantly reduces the shear mod- the Altaid belt. Mitchell et al., 1997, suggested that
ulus. This reportedly causes a larger perturbation in P- the different coda Qc magnitudes across Eurasia were
wave velocity than density, which in turn signifies proportional to the length of time elapsed since
partial melt (Halderman and Davis, 1991). the most recent episode of large scale compressional tec-
tonic activity.
They interpreted low coda Qc as resulting largely
from hydrothermal fluids generated by tectonic activity
Table 10.9 Typical Vp-Qp inter-relation, with depth as the
or heating, residing in permeable portions of the
important variable. (Derived from Iwasaki et al., 1994 data, Eurasian crust. Crustal Qc appears to increase with
reproduced in Figure 10.59) time, as fluids are lost to the surface or absorbed by
metamorphism.
Vp (km/s) 3 to 4 5.9 6.2 6.3 6.4 6.5 Companion papers that appeared in the same num-
Qp 100 150 200 300 400 500
ber of Pure and Applied Geophysics, described recently
Depth (km) 1 2 5 8 12 15–20
interpreted continental structures of Qo (the Lg coda

Figure 10.60 Simplified tectonic map of Eurasia. A tomographic plot of Qc at 1 Hz was given for each 3° by 3° cell. Mitchell et al., 1997.
Seismic quality Q and attenuation at many scales 231

at 1 Hz) for South America (DeSouza and Mitchell, Several of the above (1998) Qo ranges for regions of
1998), and North America (Baquer and Mitchell, the United States showed some differences to the earlier
1998). These authors used 389 seismic recordings in the study of Singh and Herrmann, 1983, where a broad
case of S. America, and 218 in the case of N. America, belt in the eastern region below the Great Lakes, had
to produce back-projection tomography, by inversion, values varying from 1000 to a maximum of 1300. The
giving regionalized maps of Qo and of its frequency lowest belt of Lg coda Q along the western coast
dependence (') at 1 Hz. (Oregon, Washington, California) was given as 200 in
The Lg phase was explained by Mitchell and Hwang, these earlier studies.
1987, as being prominent on regional short-period seis-
mograms, where in stable (high Q) continental regions,
it can be observed to distances as great as 4000 km, and 10.7.5 Effect of thick sediments on
forms the basis of magnitude scales for small earth- continental Lg coda
quakes, as recorded over regional (i.e. large) distances.
The main Lg phase is followed by a coda, the main Baquer and Mitchell, 1998 emphasised the role of thick
duration of which can also be used to determine the deposits of Mesozoic and younger e.g. Cretaceous sedi-
magnitude of regional events. However, the later part of ments, typically sandstones and shales, in significantly
the coda may reportedly not be coherent across arrays reducing Qo in various regions of North America, while
of seismograph stations, indicating that part of the coda older sedimentary rocks did not. They also cited the
is due to scattering. ‘positive’ effect of dolomites and limestones in main-
According to these most recent 1998 studies, the taining high Qo, and of fluids that had been lost with
seismically active South American Andean Belt was time. Earlier work by Mitchell, 1995, had suggested
typified by low Qo (250–450), in a similar manner to that seismic Q was influenced in a ‘positive’ (less atten-
the low Qo (250–300) region west of the Rocky uating) direction by the time elapsed since the most
Mountains (the Basin and Range province and active recent major episode of tectonic activity. This seemed
Californian coastal regions). In South America there to be supported by these most recent measurements,
were broad regions of very high Qo (700–1100) span- and by the Eurasian studies.
ning the central Brazilian shield, and the Amazonian In earlier studies, Mitchell and Hwang, 1987 had
and Paraná Basins, whereas in North America the high- investigated in some detail, whether the lateral vari-
est Qo region was the Northern Appalachians and some ations in Lg attenuation across the United States, could
of the central lowlands (650–750). The Gulf Coastal be explained by known variations in the thicknesses of
Plain and the southern portion of the Atlantic Coastal shallow sedimentary layers. They stated that many of
Plain had intermediate values (400–500), while the the features of the coda at frequencies near 1 Hz could
Atlantic Shield in South America also had intermediate be approximately duplicated in synthetic seismograms
values (450–700), these last possibly related to the tec- produced by plane-layer models, which included layers
tonic and igneous activity that occurred during the of low-velocity surface sediments.
break-up of Gondwanaland. As they pointed out, soft and unconsolidated sedi-
The authors of these continental studies suggested ments could be characterized by very low velocity and low
that the low Qo in the Andes, particularly in two belts Q values. However, deep sedimentary basins bounded
across the southern and northern Andes, was probably by sharp discontinuities could influence Lg by scatter-
related with higher upper mantle temperatures, or that ing (or even wave blockage, as described by Baumgardt,
there were more, deep hydrothermal fluids in these 1985), and thick sediments of low Q could cause
belts. Fluids in the upper crust, and the energy loss they rapid attenuation due to intrinsic absorption. Their
represent, were also cited as the likely reason for low Qo assumption, based on earlier studies, was that Lg Q
in the region west of the Rocky Mountains, with varia- and coda Q were approximately equal, and could
tions in that region caused by variable amounts of flu- show regional variations of greater than 1:6 across the
ids in faults, joints and rocks of variable permeability. United States.
They cited recent studies that showed that a shear veloc- Two aspects to be investigated were how far regional
ity transition from high to low velocity, lay further west variations could be explained by sediment of different
at ‘intermediate’ depths between 25 and 100 km, than age (as we have seen in the latest studies), and why there
at greater depths than this. could be low and laterally varying values of Q in the
232 Rock quality, seismic velocity, attenuation and anisotropy

depth, reproduced in Figure 10.62. The implicit ‘geo-


metric’ similarity to the velocity  depth structure of the
jointed rock mass Q-Vp-M model (Barton, 1995), is strik-
ing. Possibly the rock mass quality Qc value is about 1/10
of the coda Q, which would then give an equivalent
velocity scale along the top of Figure 10.61b, stretching
from about  3.5 km/s (due to porosity effects) to about
5.8 km/s (porosity ‘compensation’ by 2 km depth). A
rock quality Qc  1 suggests Vp  3.5 km/s (less, with
significant porosity), while a rock quality Qc  200 sug-
gests Vp  5.8 km/s (less, with significant porosity, but
(a) partly compensated by the depth effects of 2 km of
sediment). Conversion of rock quality Qc to modulus of
deformation again indicates the potential match of seis-
mic Qp and Emass, when given in units of GPa.
As Mitchell and Huang noted, there was clear evi-
dence for lower seismic Q at depths of less than 400 m.
For the modelling of sediment-layer effects, the authors
used the stepped trend in Figure 10.61a, where the fol-
lowing was suggested:
● 0–100 m Q  30
● 100–300 m Q  50
● 300–600 m Q  75
● 600 m Q  100
In their synthetic modelling the authors assumed fre-
quency independence of seismic Q below 1 Hz. The
authors found that their low seismic Q thick-sediment
models, while applicable and explanatory of Lg coda
Q in those regions with seismic Q ranging from 400 to
1300, were not applicable to the lower Q of the western
United States. Surprisingly, the authors, at that time,
did not mention jointing and faulting in (the seismic-
(b)
ally more active) western United States, as a likely cause
Figure 10.61 a) An assembly of Q ⫺1 data for sandstone and shale for lateral and depth variation of the generally low Q in
sediments as a function of depth. (See Mitchell and this region of mostly crystalline basement.
Huang, 1983 for complete references). 1: Pierre
shale, 2: Gulf Coast sediments, 3: various VSP data
sets, 4: unknown, 5: unconsolidated sediments, 10.8 Some recent attenuation
6: San Francisco Bay sediments, 7: El Centro area measurements in petroleum
sediments. b) Reversal and rotation of data by the
reservoir environments
writer to match Figure 10.62 Vp – Qc format. (See
Chapter 5 for derivation).
As will also be observed in many of the Vp, Vs, Qp, and
Qs sets of data to be presented in Chapter 13, the attenu-
crystalline crust, where deep accumulations of sedi- ation is a relatively more sensitive indicator of the degree
ment were absent, as in the western USA. Mitchell and of saturation than velocity. In particular, the Qs/Qp
Hwang assembled Q seis values for sandstones and shales, ratio in sedimentary rocks is much more sensitive to the
excluding limestones due to their high Q seis values. degree of saturation than the Vp/Vs velocity ratio.
Figure 10.61a shows this sedimentary Q⫺1 data, with Seismic Q has therefore become increasingly important
reversal and rotation by the writer (Figure 10.61b) to in hydrocarbon exploration. For similar reasons,
match the plotting format of rock mass quality Qc versus Winkler and Nur, 1979, suggested using these relatively
Seismic quality Q and attenuation at many scales 233

Figure 10.62 Rock quality Qc-Vp depth model for comparison to previous data set. Making a gross approximation of rock quality Qc 1/10th
of Q, an approximate velocity scale of 3.5 km/s to 5.8 km/s (less, due to sediment porosities) is suggested in Figure 10.61b.

large changes in seismic Q with saturation as an earth-


quake predictor, since pre-earthquake dilatancy could
affect the degree of saturation in fault zone rocks. These
saturation/partial saturation effects were reviewed in
section 10.2, but derived only from intact bars of
dynamically excited sandstone,
Anomalously low seismic Q values at depth, adjacent
to a 5 km deep well where VSP was performed, were
reported by Keehn and Kanasewich, 1987. The spectral
ratio method was used to obtain values of seismic Q of
10 in a Lower Triassic sandstone, and a second low seis-
mic Q of 12 at about 4,000 m depth. (Figure 10.63). The
authors considered that they were observing the effects of
scattering caused by intrabed multiples, together with
intrinsic attenuation associated with the sandstone
lithology.
The strongly attenuating zone between 1,930 and
2,320 m depth was associated with an almost uninter-
rupted Lower Triassic sandstone layer. The authors
Figure 10.63 Depth plot of 1/Qseis and Q, smoothed with a
referred to studies showing that increased content of sand,
200 m running average, together with the interval
rather than shale, were responsible for high attenuation. P-wave velocities. The well is on Melville Island, in
For some reason, the possibility of a fractured reservoir in the Canadian Arctic. Keehn and Kanasewich, 1987.
this location was not directly referred to, possibly due to
confidentiality. remainder of the hole, pores with smaller aspect ratios,
The likelihood that oil bearing rock in a jointed more pore space, and a higher degree of saturation’.
reservoir was being described, is however evident from Cross-well tomography data of Quan and Harris,
their conclusions. ‘The existence of the two low-Q 1997 shown in Figure 10.64, emphasises the generally
zones may be attributed to one or more of the follow- high level of attenuation in transversely isotropic (‘layer
ing factors which contribute to high attenuation: the cake’) sedimentary series. Seismic Q values were between
presence of fluids of higher viscosity than those in the about 20 and 90. The attenuation coefficient () in the
234 Rock quality, seismic velocity, attenuation and anisotropy

central log, is given by the relation Qseis  f/  V, employed, as emphasised by the important set of
where V is the wave speed, and f is the frequency. It is attenuation data from the Imperial College test site in
apparent from the figure that there were certain simi- NE England, reproduced earlier, in Figures 10.34 and
larities, in this case, between the cross-well velocity 10.35, from Sams et al., 1997. The differences in attenu-
structure and the seismic quality Q. The 8 to 17 K ation between sonic, cross-well, VSP and ultrasonic
ft/sec velocity scale converts to 2410 m/s – 5120 m/s. measurements in the same formation were significant
Each of these values (Vp and Qseis) showed a clear differ- The four boreholes utilised by Sams et al., 1997,
entiation between shale, limestone and clay, and there were drilled to about half the depth of the above, to just
was a certain indication of a less jointed (or less porous) a few hundred metres depth in a layered sequence of
area in the overlying chalk. limestones, sandstones, siltstones and mudstones. The
The cross-well Q seismic data shown in Figure 10.64 variability of 1/Q for each type of seismic survey was
are unfortunately specific only to the frequency range due both to rock heterogeneity (i.e. implicit rock qual-
ity Q-value variations) and frequency variations.
The authors Hustedt and Clark, 1999, drew atten-
tion to the fact that the seismic attenuation factor Q is
an important parameter in the processing and interpret-
ation of seismic data, both because of the detrimental
effect it has on the data, and because it can itself be an
indicator of rock properties.
Hustedt and Clark referred to the QVO (Q versus off-
set) technique that had recently been introduced by
Dasgupta and Clark, 1998. This could be used for extract-
ing Q from routine marine surface seismic reflection
data. As they explained, in exploration analyst jargon:
‘The QVO method applies the well-known spectral-
ratio method to a true-relative-spectrum-processed,
Figure 10.64 Sonic log, cross-well Vp, and cross-well attenuation NMO-corrected, CMP gather.’
and seismic Q from roughly 600–900 m depth, in They compiled QVO-derived seismic Q-values in rela-
the BP Devine Test Site (Quan and Harris, 1997), tion to interval velocities, from a variety of hydrocarbon
reviewed by Pride et al., 2003. Centre frequency was exploration settings, as shown in Figure 10.65. Q-values
1750 Hz. ranged from 50 to 700–800, suggesting that some of

Figure 10.65 A compilation of QVO (Q versus offset) data for seismic quality Q versus interval velocity. Hustedt and Clark, 1999.
Seismic quality Q and attenuation at many scales 235

the ‘rules-of-thumb’ attempting to relate an approxi-


mate Q to the interval velocity, may be inappropriate.
The potential ‘central trend’ of the data – a steepening
curve – was unfortunately defeated by a dense ‘clump’ of
low seismic Q (40 to 50), yet displaying medium high
velocity, of about 3.9 to 4.1 km/s.
In the context of jointed rock (possibly inapplicable
at the undefined, but presumed several kilometres
depths), a certain Vp – Q coupling, as in Figure 10.64,
would be understandable. In relation to the ‘curving
trend’ of much of the data in Figure 10.65, a slightly
better fit than the straight lines could be Qp  (Vp-
500)0.6, implying the ‘disappearance’ of a measurable
Qp at Vp as low as 0.5 km/s.

10.8.1 Anomalous values of seismic


Q in reservoirs due to major
structures

Dasgupta and Clark, 1998, reported seismic Q values


of 46 and 130 for top chalk and base chalk, from some (a)
North Sea data. In Figure 10.66 Qp is shown falling
from 100 to a minimum value of about 50 in the anti-
clinal crest. The implied reservoir values (13 to 33)
caused by gas effects agreed with trends shown in labo-
ratory studies (See Chapter 13).
A low value of seismic Q for a fault zone (or possibly
several faults) encountered in a well in the North Sea
(Qseis averaging 45), was described by Worthington and
Hudson, 2000. Figure 10.67 shows the roughly 1000 m
to 2000 m depth trace of the fault in the seismic
migrated time section, reproduced from Harris et al.,
1997, and its approximate intersection with the well. The

(b)

Figure 10.66 Qp values interpreted in a North Sea UK sector anti- Figure 10.67 Anomalously low seismic Q related with fault zones
clinal crest, with further reduced values in a gas bear- in a North Sea reservoir. Worthington and Hudson,
ing pay zone. (Dasgupta and Clark, 1998). 2000, from Harris et al., 1997 data.
236 Rock quality, seismic velocity, attenuation and anisotropy

fault zone caused an abrupt increase in attenuation, rela- due to the fracturing, which indicated permeabilities of
tive to the Triassic and Lower Jurassic age sandstones, 2.5 to 5 milli-darcies.
siltstones and claystones that were predominantly The en echelon vertical fractures were very short 0.4
encountered in the well. to 8 cm, frequently occurring features, perpendicular to
Worthington and Hudson described their modelling the bedding in the Brown Shale. The joint-like features
of the effects of a down-going P-wave between 1000 and that were more dominating in the Antelope Shale, were
2000 m depth, by assuming that a fault or several faults, nearly vertical, and also had modest heights averaging
intersected the transmission path. By using a compli- only 13 cm, due to the bedding thickness limits. Joint
ance model of a major discontinuity with not com- densities were from 0 to 2 per meter. There were also
pletely conforming opposite faces, they showed the need less frequent larger fractures and micro-faults, also per-
for a remarkable, but actually very realistic, inequality of pendicular to the bedding.
the normal and shear compliances. The cross-well tomography, with sources and
We will examine these important parameters (whose receivers at 1.5 m (5 feet) intervals over 457 m (1500
inverse is dynamic stiffness), in detail in Chapters 15 feet) of the reservoir formation, showed expected, dis-
and 16, seeing the similarity of their inverted magni- tinctly layered velocity trends in the range of about 2.4
tudes, to the normal stiffnesses of joints, clay-filled dis- to 3.4 km/s over a selected depth range of 3,900 ft
continuities, or faults that are more familiar in the (1,190 m) to 4,600 ft (1,400 m).
macro-displacement world of rock mechanics. This Figure 10.68 indicates the P-wave velocity and com-
subject was also addressed earlier in this chapter, con- puted Qseis for an interval in the Antelope Shale. Also
cerning the important work of Pyrak-Nolte et al., 1990 shown is the core-plug permeability. It is particularly
related to dynamic and static loading tests on joints. interesting to note the good ‘geometric correlation’ of
Vp and Qseis in the fractured part of the formation with
lower core-plug permeability. This was the Brown
10.8.2 Evidence for fracturing Shale, which contained both styles of fracturing. In the
effects in reservoirs on lower parts of the formation with more sand, there was
seismic Q a marked increase in core-plug permeability, but Qseis
remained low, even below 20, probably due to intrinsic
Evidence for the subtle effects on seismic Q, of fractur- squirt-flow attenuation in the sand and carbonate beds,
ing in petroleum reservoirs, was given by Parra et al., where the P-wave velocity was markedly higher. The Q
2002, who described field characterization at the Buena seismic data is reproduced at more exaggerated scale in
Vista Hills reservoir, in California. They described the Figure 10.69. This shows the strong influence of the
use of seismic Q derived from high-resolution cross- two styles of jointing that were described.
well seismic data, to detect vertical, joint-like tectonic The authors also conducted poro-elastic numerical
fracturing dominating in the Antelope Shale and en modelling, based on a Biot squirt flow attenuation
echelon, sigmoidal, vein fracturing that was restricted mechanism. They demonstrate in Figure 10.70 the effect
to the Brown Shale, where the joint-like fractures also of frequency and azimuth angles on computed attenua-
occurred, but with less frequency. The Brown Shale dis- tions, referring to the actual vertical fracture azimuths of
played both low Qseis, and low P-wave velocity. 0° to 30°.
The sand-shale sequences were too finely layered to It is interesting to note from their modelling of the
be detected by sonic logging, with layer thicknesses and Brown Shale, that there was little attenuation of seismic
bed thicknesses ranging from fractions of centimetres waves propagating from sources at the surface, where Qseis
to tens of centimetres to meters. The Antelope Shale was a surprising 1000. In contrast, they found that the en
formation, containing the highest densities of jointing, echelon and joint-like features in this shale were strongly
consisted of thin, siliceous, clay-free shale beds, with attenuating to seismic waves propagating parallel to strat-
intercalated thin laminae of clayey sand and carbonates. ification, and perpendicular to the fractures (Qseis of 20),
A 290 m core interval contained nearly 750 sand lam- in the frequency range of sonic and cross-well seismic.
inæ. The cores averaged 28% porosity, but had only The lower Antelope Shale with its frequent sand
70 micro-darcies permeability due to the dominance of beds, indicated a higher attenuation of seismic waves
siliceous shale. Since only 5% of the rock consisted of propagating from the surface, with Qseis typically 100,
the sands, hydrocarbon production was assumed to be while waves propagating parallel to stratification and
Seismic quality Q and attenuation at many scales 237

(a)

Figure 10.68 Vp and Qseis results computed from velocity dispersion


data (1 to 10 kHz), for a selected 1,220 to 1,330 m
deep section of the en echelon fractured and infre-
quently jointed Brown Shale, and the lower, well
jointed, sand-laminae bearing Antelope Shale, Buena (b)
Vista Hills field, California. Also shown is the core-plug
permeability, which is highest where sand is more fre- Figure 10.69 Low seismic Q in jointed Antelope shale, and in en
quent in the Antelope Shale. The uppermost low Qseis echelon fractured Brown shale, plotted at exaggerated
zone corresponds to the Brown Shale. Parra et al., 2002. scale. Buena Vista Hills studies described by Parra
et al., 2002.
perpendicular to ‘open’ fractures, indicated seismic Q
values of 18 to 20. For their modelling, they had et al., 2005 described the use of attenuation and vel-
selected permeabilities perpendicular to the fracture ocity tomography, using an array of ocean bottom seis-
systems of 50
d, similar to the matrix, while parallel to mometers (OBS), crossed by a dense pattern of shot
the fractures they selected 5 md. (Note: 1 darcy  108 lines, on the western continental margin of Svalbard,
cm2, or 101 2 m2, which is roughly equivalent to an on the lower part of the continental slope, close to an
engineering unit of 105 m/s for water). active mid-ocean ridge.
The uses of seismic Q appear to be expanding rap- Data was acquired within the EU Hydratech Project.
idly, as time goes by, due to its greater sensitivity to An important boundary condition was a reflector,
some physical properties than seismic velocity. Rossi marking the boundary between gas-hydrate and free-gas
238 Rock quality, seismic velocity, attenuation and anisotropy

10.8.3 Different methods of analysis


give different seismic Q

On several occasions, different methods of estimating


seismic Q have been compared by the same authors. We
will briefly review two such cases, in order to emphasise
both the difficulties that can sometimes arise, and the
potential errors involved. As an observer of geophysical
results, rather than a practitioner, it is not possible to
judge whether the levels of error reported here, are a
serious threat to the use that the geophysics community
is currently making of seismic Q, in all its various forms.
Badri and Mooney, 1987 used several processing
methods, in both time and frequency domains, to com-
pute the seismic quality factor Q for water-saturated,
Figure 10.70 They computed Qseis values of 15, 19 and 28 in the
unconsolidated sediments. The methods used included
frequency range of 120 to 1000 Hz using an measurements of the spectral amplitude ratio, peak-to-
assumed squirt flow length of 3 cm for the modelled peak and first-peak amplitude ratio, rise time, pulse
fractures. (Parra et al., 2002). broadening, and the Futterman causal attenuation opera-
tor for attenuating signals.
bearing sediment zones below the reflector. Qp showed The authors used compressional seismic waves gener-
a strong decrease from around 200 to values declining ated from explosive sources ranging in size from 1 to
below 50 and even below 25, below this reflector, corres- 64 mg of silver azide, at a depth of 7.6 m below the 70%
ponding to the sediments containing free-gas. The high saturated, silty sandy clays, near Wendover, Utah. The
values of Qp corresponded to the expected gas-hydrate hydrophone receivers were spirally distributed at
zone, and its probable ‘strengthening of the solid frame’ distances ranging from 25 to 200 m from the source. The
of the rock. The authors generally found good spatial computed seismic Q values showed a remarkable vari-
agreement between the Qp and Vp variations, both ver- ation, as indicated below. The authors suggested that the
tically and laterally. The corresponding Vp values were spectral amplitude ratio method was probably the most
about 1.75 to 1.8 km/s above the reflector in the gas- reliable, as it is applicable independent of the source.
hydrate zone, declining rapidly to 1.5–1.6 km/s in the
1. Spectral amplitude ratio method, with five explo-
free-gas zones below.
sive source sizes, over frequencies of 450–725 Hz.
The authors also noted that the small graben-causing
Average Q  23.
faults in the area, correlated with the Qp and Vp reduc-
2. Peak amplitude ratio method. Average Q  123.
tions. Interestingly, the Qp values began to decline just
3. Rise-time method. Range of Q  50–207.
above the reflecting boundary, where Vp was still high.
4. Pulse broadening technique. a) Quarter-cycle
The authors considered that this might have been due to
measurement: range of Q  25–158. b) Half-cycle
interference in the frequency-shift calculations, caused
measurement: range of Q  26–114.
by the liquid to gas phase change. One other anomalous
5. Futterman causal attenuation operator. Range of
result when comparing the two parameters was the
Q  200–300.
almost full recovery of Vp at greater depth, while Qp
increased only to about 75. Intuitively, in view of the unconsolidated material and
At greatest depth there was a further unexplained Qp of shallow depth involved, one could be permitted to assume
about 20, while the velocity attained its highest value of that the three lowest seismic Q values listed above (23,
1.9 km/s. The consistent rise in velocity in the deeper lev- 25 and 26), were likely to be the most realistic, and the
els resembled typical velocity-depth (stress-closing-struc- author’s preference for the spectral amplitude ratio
ture) trends seen in particular in Chapter 11. Possibly a method, seen in much of the literature, seems likely to
strongly developed structure, under significant effec- be the most reliable.
tive stress, would be capable of showing an ‘increased Toverud and Ursin, 2005 went further than the
Vp, reduced Qp’ reaction. above, and compared eight methods of determining
Seismic quality Q and attenuation at many scales 239

seismic Q, using (almost) zero-offset VSP for three sep- was 427 m of mainly shales, with some limestone and
arate zones, obtained from a well off the Norwegian marls; 3335–3650 m was 315 m of marl/limestone,
coast. The source was deployed at 4 m depth at 40 m of with some shale; 3650–3907 m was 257 m of sand-
horizontal offset from the well. Depths analysed ranged stone, siltstone and shale).
from 2907–3907 m, using 10 m intervals. A minimum The results for the second layer (marl/limestone with
frequency close to zero up to 90 Hz was indicated. some shale) gave considerably higher Q and more vari-
The authors evaluated eight different attenuation ation. This zone had the highest content of limestone and
models, using a least squares model-fitting approach. marl, which perhaps explains both the higher values
They used the geometric ray approximation approach (Q  90–160), and the large range. The apparent good
of Ursin and Arntsen, 1985, for point source, vertical wave correspondence of the models for layers 1 and 3 was a
propagation in a 1D viscoelastic medium, with plane- function of the use of the minimum (normalized) misfit in
wave reflection coefficients. A formula for the complex Table 10.10. An example of their normalized misfit analy-
velocity was assumed, with inversion of the attenuation sis, for the Kolsky-Futterman model, is shown in Figure
parameters at three different depth intervals, to obtain the 10.71. They considered that this performed slightly better
parameters in three homogeneous layers: (2907–3335 m than the other models, except in the middle layer.
In reality, and in Figure 10.72, the authors show that
Table 10.10 Comparison of eight methods for estimating seismic there is actually a lot of difference between the models
Q. Toverud and Ursin, 2005. Minimum normalized concerning their frequency-dependence. As may be
misfits are shown. (For brevity, results for two of the noted, the Kolsky-Futterman and Kjartansson models
three layers are selected here. Velocities have been assume almost negligible frequency dependence.
rounded to the nearest 10 m/s. Variations were less
than 5 m/s).

Model Layer Qp Vp m/s

Kolsky-Futterman 1 32 3.14
3 36 2.98
Power law 1 35 3.14
3 39 2.98
Kjartansson 1 31 3.14
3 37 2.98
Müller 1 34 3.15
3 40 2.98
Azimi’s second law 1 36 3.15
3 44 2.98
Azimi’s third law 1 34 3.14
3 40 2.98
Cole-Cole 1 22 3.14
3 25.5 2.99 Figure 10.72 A comparison of the eight models for interpreting
Standard linear solid 1 27.5 3.14 seismic Q, for the case of layer 1. Large differences
(SLS) 3 32 2.99 between models are indicated concerning frequency
dependence. Toverud and Ursin, 2005.

Figure 10.71 An example of Toverud and Ursin, 2005, normalized misfit analysis, using the Kolsky-Futterman model, for the three layers
of VSP data analysed for seismic Q.
11 Velocity structure of the
earth’s crust

This chapter summarises the velocity structure of the Figure 11.1, reproduced from Kearey and Vine,
continental crust, the continental margins, and the sub- 1996, shows a ‘familiar’ increase in Vp and Vs through
ocean spreading ridges, where zero-age crust is forming. the crust and upper and lower mantle. However, due to
The velocity-depth models of large scale (100 m to 1 km the great pressure and much higher densities of the rocks
to 50 km) naturally represent an extension of near-surface involved, the magnitude scale for Vs (4 to 7 km/s) now
experiences from Part I. However there are some impor- looks more ‘familiar’ than the exceptional 8 to 13 km/s
tant parallels and points of basic similarity, especially range of the P-wave velocities. The depths of nearly
beneath 3 km of ocean. Although mainly concerned with 3000 km of course exceed by up to five orders of magni-
velocity-depth trends and their reasons, there are also tude, the near-surface phenomena reviewed in Part I.
some parallel seismic Q results, where separation into When reaching the outer core and assumed fluid con-
Chapter 10 would have been undesirable. ditions beyond 3000 km, the S-wave is shown as falling
to zero due to the loss of shear strength, and the P-wave
also shows a dramatic fall of some 5 km/s to ‘only’ 8 km/s,
11.1 An introduction to crustal increasing thereafter to about 10 km/s, before reaching
velocity structures the inner core with its increased density, where the S-wave
makes a return, albeit to less than 4 km/s.
The text of Part I of this book was dominated by civil In 1909, Mohorovicic interpreted a first arrival P-wave
engineering scale velocity-depth and rock quality trends. of magnitude 5.7 km/s within 200 km of an earthquake
In Part II, Chapter 11 we will now concentrate on the epicentre, and another first arrival P-wave of 7.9 km/s
velocity-depth trends of the continental and oceanic at greater distance from the epicentre, as evidence for a
crusts. However, when contemplating assembling a velocity discontinuity. This is now termed the Moho.
review of the velocity structure of the crust, a complete Figures 11.2a, b show travel time-distance gradients
section to the inner core of the earth, as summarised in consistent with the Moho velocity discontinuity both for
Figure 11.1, is clearly an important starting point for the case of the thicker sub-continental crust (shown here
delineating the outer boundaries of behaviour. as 54 km) and for the case of the much thinner sub-ocean
crust (shown here as about 12 km). The direct wave (Pg)
and the refracted wave (Pn) show different gradients.
At first, there was suspicion of another velocity dis-
continuity at intermediate (sub-ocean) depth, based on
Conrad, 1925. A velocity increase from 5.6 to 6.3 km/s
is shown in Figure 11.2b. It is now known that the
Conrad discontinuity is not always present and a grad-
ational increase with depth is generally seen. We shall
see much evidence of these gradational increases in
velocity in this chapter.
Information about the uppermost parts of the earth’s
crust is now available from direct sampling in ultra-
deep boreholes, and indirectly from experimental data
Figure 11.1 A simplified velocity structure through the crust and on velocities measured over ranges of temperature and
mantle down to the inner core. Kearey and Vine, 1996, pressure consistent with crustal conditions. Pressure
from Hart et al., 1977. (total stress) increases at a rate of about 30 MPa per
242 Rock quality, seismic velocity, attenuation and anisotropy

Table 11.1 A simplified, classic model for the seismic structure of


oceanic crust (from Bott, 1982, reproduced by Kearey
and Vine, 1996).

P-wave velocity Average thickness


(km/s) (km)

Water 1.5 4.5


Layer 1 1.6–2.5 0.4
(sediment)
Layer 2 3.4–6.2 1.4
Layer 3 6.4–7.0 5.0
Moho discontinuity
Upper mantle 7.4–8.6

gabbroic anorthositic rocks which therefore give a higher


range of velocity (Kearey and Vine, 1996).
The oceanic crust of the earth is much thinner than the
continental crust, and is usually about 6 to 7 km thick,
beneath an average water depth of 4.5 km. Table 11.1
reproduced from Kearey and Vine, 1996, and based on
Bott, 1982, gives velocities for a simplified layered model,
while Figure 11.3 shows more detail and a more gradual
velocity increase (dash line) based on inversion techniques.
The type of measurements traditionally required to
obtain such information, so-called reversed deep-sea
refraction, are illustrated in Figure 11.4. Here we can see
Figure 11.2 a, b Diagrams of time-distance gradients that demon- the typical layer 1, 2, 3 and Moho separations of velocity,
strate the sub-continental and sub-ocean evidence for that were state-of the-art prior to more extensive investi-
the Moho. From Kearey and Vine, 1996. gations (and investments), from the end of the 1970s.
Measurements of the type illustrated were extended
from continental shelf to deep sea, giving as in Figure 11.5,
kilometre due to the high average density, but there a good illustration of the relative thickness of continental
is an initial pore pressure increase to perhaps 20 km and ocean crust, which are due to great differences in
or more. At shallower depth, it is common to assume age, as we shall see later in this chapter.
about 16 or 17 MPa/km increase of the effective stress Of course there are also anomalies to complicate the
(i.e. 26 minus 10  16 MPa/km) in crystalline rock, simple picture of increasing velocity (and seismic Q)
and closer to 10 or 12 MPa/km in less dense hydrocar- with depth, as implied so far. A seismic low-velocity
bon reservoir sediments, neglecting over-pressured zones. zone at depth in the crust is widely accepted as evidence
Temperature increases at a rate of about 25°C per kilo- of a region of partially molten rock. It can explain the
metre up to the Moho, which usually varies from 20 to occurrence of low seismic Q and large negative gradi-
80 km depth beneath continents. ents of both velocity and seismic Q with depth. It was
The uppermost 5 km of the crust shows a rapid increase also considered by Mavko and Nur, 1975 as a likely
in deformation modulus and density, as pore space and zone of relaxation that could be responsible for tran-
joints are closed. However, the thermal expansion partly sient deformation following large earthquakes on plate
balances the increase in seismic velocity, and P-wave boundaries. Their ‘melt squirt’ mechanism – possibly
velocities above about 6.5 km/s do not appear to be involving flow of molten rock between cracks of differ-
common. The P-wave velocity range in the lower crust, ent orientation to the changed stress field, apparently
from about 6.5 to 7.6 km/s is explained by chemical gives a relaxation time of the right magnitude (a few
transformation to more dense phases, e.g. basalt to garnet years) to explain transient deformations that may follow
granulite to ecologite, or by the presence of higher density the ‘elastic rebound’ phase of deformation, following
Velocity structure of the earth’s crust 243

Figure 11.3 Simplified models, dating from 1965 and 1978, of supposed layering in the oceanic crust. On the left, with the benefit of
improved inversion techniques, is a Spudich and Orcutt, 1980, and Harrison and Bonatti, 1981, interpretation of a more gra-
dational increase in velocity with depth.

was complex, the seismic structure was simple. The


average velocity-depth gradients for the investigated
terrain were as shown in Figure 11.7. Here the field
data is compared with relevant laboratory data.
To conclude this introduction to crustal velocities we
will return to greater depth, by first considering the vel-
ocity and seismic Q structure within and above descend-
ing crustal material, followed by a glimse of the deeper
velocity and seismic Q trends. One of the most typical
subduction zones in the world is the north-eastern Japan
arc. The oceanic Pacific plate subducts downwards into
the mantle at a convergence rate of about 10 cm/yr and at
an angle of 30° and steeper at greater depth. Many shallow
earthquakes occur beneath the Pacific ocean along the
upper boundary of the Pacific plate. Intermediate-depth
Figure 11.4 Reversed deep-sea refraction, using two ships and explo- and deep earthquakes are generated within the subducted
sive charges. From Bott, 1982, based on Talwani, 1964. Pacific plate. Beneath Japan, shallow earthquakes also
occur in the upper crust of the continental plate. Active
the accumulation of surface deformations that may be volcanoes are distributed on the land area, parallel to
approximately ‘cancelled’ during a large earthquake. the trench axis.
A multidisciplinary investigation of the tectono- A modern interpretation of this north-eastern Japan
stratigraphic terrain that compose the Alaskan litho- convergent margin is shown in Figure 11.8 from
sphere by Beaudoin et al., 1992a, revealed low-velocity Hasagawa et al., 1994. In the base of the mantle wedge,
(6.4 km/s) rocks extending to a depth of approximately low-velocity, low seismic Q zones are distributed in par-
27 km. In this case, little complexity was suggested, with allel to the dip of the high seismic velocity, high seismic
seismic layering typically as shown in Figure 11.6 Q subducting plate. Decompression melting within the
The bedrock composition was metasediments, meta- ascending flow of hot mantle material from depth pro-
granitic rocks and granitic plutons. Principal miner- duces low seismic velocities and high seismic attenuations.
alogical compositions were quartz, plagioclase and The lower portion of the crust and mantle wedge are
mica, which reportedly have similar average compres- governed by creep or flow, and are weak and incapable
sional wave velocities. Therefore although the geology of supporting high stress. According to the review by
244 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.5 Refraction lines from Argentina and 1000 km eastwards into the Atlantic Ocean. From Ewing, 1965, reproduced in Bott, 1982.

Figure 11.6 Typical ray diagram (for shotpoint 52) for the East Central Alaska crust (Beaudoin et al., 1992a).

Hasagawa et al., 1994, horizontal compressive stress the depths of 100 km and 1200 km. These extreme
caused by the convergence between the subduction plate depth trends for Vp and Q p are shown in Figure 11.9.
and the overlying continental plate is supported mostly Strong variations in the upper 20 km of solid crust are
by the upper 15 km of the crust. This is a strong seismic suggested in this large-scale data, within the Q p range
zone, resulting in shallow, thrust-fault earthquakes. Stress of about 130 to 1000. There is an inferred fall of Q p
concentration will also arise beneath the volcanic regions from about 1000 to 150 through the Moho, between
where the seismogenic zone is locally thin. depths of about 40 and 80 km, followed by a rapid
P-wave velocities plotted on a depth scale of 0 to increase of Q p to 200 km depth, and a slower increase
1200 km for the western USA, determined from spectral of Q p values to about 8000 at 1200 km depth.
amplitudes of seismic body waves, given by Archambeau
et al., 1969, appear to ‘start’ at about 6 km/s rapidly
reaching about 8 km/s through the crust with intermit- 11.2 The continental velocity
tent increases to almost 12 km/s at 1200 km depth. Their structures
studies also suggested a frequency-dependent value of
seismic Q, with magnitude increasing with frequency. A definitive, updated summary of the seismic velocity
Seismic Q values inferred from the ‘anelastic dissipa- structure and composition of the earth’s continental crust
tion’ of compressional body waves and surface waves has been provided by Christensen and Mooney, 1995,
are shown to increase from about 150 to 8000 between who gave a global review based on 560 determinations by
Velocity structure of the earth’s crust 245

Figure 11.7 Average Vp-depth data (stepped line) compared with


temperature corrected laboratory data (curved lines).
Beaudoin et al., 1992a.

Figure 11.8 Schematic cross-section of crust and upper mantle in the


NE Japan convergent margin. Numerous open circles
show focal mechanisms. Solid circles show low fre- Figure 11.9 a, b Ultra-deep Vp and Q p structure interpreted for
quency micro-earthquakes in low velocity (and low seis- western USA, by Archambeau et al., 1969.
mic Q) zones beneath volcanoes. Hasagawa et al., 1994.
The average crustal thickness, weighted to correctly
more than 100 investigators. The geographic locations are represent the total global areas of each major crustal
illustrated in Figure 11.10. The data reviewed and selected type is 41.1 km, while the thinnest is in Ethiopia (Afar
by these authors covers the years 1950 to 1993. In the Triangle: 16 km), and the thickest is in China (Tibetan
refraction seismic methods applied, the apparent seismic Plateau: 72 km). The average compressional wave vel-
velocities are directly measured, while the depths of the ocity is 6.45 km/s. By chance, this is close to the ‘focal
refracting horizons are calculated from the uppermost point’ in Figure 5.36, satisfying the intact, massive rock
layer down to the deepest layer. The depth determinations quality Q-value  1000 ‘limit’ of 6.5 km/s, for an
generally have larger percent errors than the velocities. undefined, average mineralogy.
246 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.10 Locations of 560 seismic velocity-depth measurements. Christensen and Mooney, 1995.

Average compressional wave velocities of common tight structure, probably with Q-values of rock mass
crustal rocks show excellent correlations with density. quality (Barton et al., 1974; Barton, 2002) of 250–500,
Based on tests of 3000 cores of igneous and metamorphic combined with the effect of exceptionally high confining
rocks, taken to 1 GPa confining pressure (similar to 35 km stress.
depth), Christensen and Mooney estimated a mean If sedimentary rocks were included, we would likely
2830 kg/m3 density for the continental crust. The seismic be operating with a rock mass quality Q c value (Q c 
velocity data was considered to be accurate to 3%, or Q  c/100 plus porosity adjustment) less than the
about 0.2 km/s, while the depths were considered to above, depending on the effects of c 100 MPa and
be accurate to 10%. porosity  5% on the seismic velocity (see Figures 5.36
Figures 11.11a, b and c show the very clear trends and 5.37 in Chapter 5).
of crustal depth, average crustal velocity and upper In Table 11.2, mean velocities for five principal tec-
mantle velocity (the normal P-wave termed Pn by tonic provinces as a function of depths of 5, 10, 15, 20
tectonophysicists). and 25 km are given. The uppermost (lowest Vp values)
These three worldwide compilations suggest a 10 to reflect a great range of lithologies, and presumably some
20% thicker continental crust than previous estimates residual (i.e. tightly closed) jointing.
(due to under-representation of shields, platforms and The velocity-depth gradients for these five tectonic
orogens while the average velocities lie within previous provinces, and for an average continental crustal model,
Eurasian, North American and global estimates of 6.30 are compared in Figures 11.13a and b. The almost linear
to 6.55 km/s (for Vp, continental crust) and 7.7 to gradient between 5 and 25 km for the average crust dis-
8.6 km/s (for Pn of uppermost mantle). plays a gradient of about 0.6/20  0.03 s1, while the
Histograms of average velocities for 5, 10, 15, 20 and gradient between 5 and 10 km for the five tectonic
25 km depths are reproduced from Christensen and regions is approximately 0.5/5  0.1 s1. The reduced
Mooney, 1995, in Figure 11.12. gradient at greater depth is due to the expansion effect
Shallow crustal velocities of less than 5.0 km/s, caused by increased temperature.
corresponding to sedimentary rocks have not been Based on a very extensive (3000 cores) laboratory
included. (This applies to the upper 10 km.) The his- study, stretching over some ten years, Christensen and
tograms at 5 and 10 km are sharply peaked at 6.0 to Mooney, 1995, were able to distinguish anisotropic
6.2 km/s, typical of crystalline upper crust. Possibly a (mineral/fabric orientation related) velocities for a wide
minimum of ‘effective’ jointing is found at these high range of crustal rock types. The results for a confining
pressures of 140 to 280 MPa. In other words in engineer- pressure of 1 GPa (35 km depth) are reproduced in
ing terminology, we would be talking of a very massive Figure 11.14. However the authors pointed out that it
Velocity structure of the earth’s crust 247

Figure 11.11 a) Histogram of crustal thickness, from 560 meas-


urements. b) Histogram of average continental crust
velocity (Vp). c) Histogram of uppermost mantle
velocity beneath the continental crust (Pn).
Christensen and Mooney, 1995.

was not possible to take into account the effects of


larger scale anisotropy in their crustal averages, since
the presence of crustal anisotropy had only recently been
documented.
They expected maximum anisotropy in upper crustal Figure 11.12 Histograms of crustal velocity at 5 km depth intervals.
metamorphic rock with abundant phyllite and higher Christensen and Mooney, 1995.
248 Rock quality, seismic velocity, attenuation and anisotropy

Table 11.2 Velocities for five principal tectonic provinces. Christensen and Mooney, 1995.

Crustal property Orogens Shields and platforms Continental Arcs Rifts Extended crust Average crust

Vp at 5 km 5.69  0.67 5.68  0.81 5.80  0.34 5.64  0.64 5.59  0.88 5.95  0.73
Vp at 10 km 6.06  0.39 6.10  0.40 6.17  0.34 6.05  0.18 6.02  0.45 6.21  0.27
Vp at 15 km 6.22  0.32 6.32  0.26 6.38  0.33 6.29  0.19 6.31  0.32 6.31  0.27
Vp at 20 km 6.38  0.34 6.38  0.26 6.55  0.28 6.51  0.23 6.53  0.34 6.47  0.28
Vp at 25 km 5.53  0.39 6.53  0.27 6.65  0.28 6.72  0.35 6.69  0.30 6.64  0.29

grade slate, and in deeper crustal sections of amphibo-


lite and mica schist. To this one could perhaps add the
possibility of azimuthal velocity anisotropy, even at
depth, resulting from regions of strong horizontal stress
anisotropy, in e.g. thrust belts.
In general, these results are based on rock cores taken
in three mutually perpendicular directions. The authors’
reported that the change in anisotropy with depth was
minimal for most of the rock types. Anisotropies reach
9.5%, 13.0% and 17.2% on average for the mica quartz
schist, phyllite and slate respectively. Even at upper
Figure 11.13 a) Average velocity depth trends for five tectonic mantle depths, azimuthal-dependent Pn velocities are
provinces, compared to b) the average crust. consistently shown, particularly along (and across) the
Christensen and Mooney, 1995. axes of continental rift structures. When one considers
the added effect of near-surface (upper 5 km) jointing
that may be parallel or sub-parallel to dominant fabric,
anisotropy will presumably tend to increase on average
as the surface is approached.
Average velocities at 20 km depth equivalent and at
309°C (using average heat flow assumptions) for each
rock type are shown in Figure 11.15. The majority of
rocks lie between velocities of 6.0 and 7.0 km/s.
As has been noted, the effect on velocity of increased
heating in a given laboratory test sample tends to coun-
teract the effect of increased pressure due to thermal
expansion effects. However a single rock type ‘taken from
5 km to 50 km’ does not show the average continental
crust velocity-depth gradient, which is brought about
by a combination of lithological changes, mineralogical
changes and temperature-induced expansion.
Within the mid-crustal depths of 10 to 25 km, where
amphibolite facies rocks are likely to comprise the bulk
of the crust, there is a gradual change in composition
from granitic gneiss and tonalitic gneiss to mafic min-
eral assemblages rich in amphibolite. At greater depth,
the garnet content increases. These gradual changes give
the crust its composite average velocity-depth gradient,
Figure 11.14 Average anisotropies for laboratory samples at in relation to the single rock type, laboratory trends
1 GPa confinement. Christensen and Mooney, 1995. shown in Figure 11.16.
Phyllite, slate and schist dominate as expected, even The contrasting trends shown in Figure 11.16 are
at these high pressures. again from Christensen and Mooney, 1995, who must
Velocity structure of the earth’s crust 249

There were 2592 tiles of 5°  5°, and more than


2000 available sets of field measurements of oceanic
and continental crust. Primary continental and oceanic
crustal types and mean Vp as a function of depth are
shown in Figures 11.18a and b. Note the predominance
of average velocities 6.0, 6.1 and 6.2 km/s for the upper
crust (range 5.7 to 6.3 km/s) and mostly 6.6 km/s for
middle crust (range 6.4 to 6.7 km/s).
On continents, the P-wave velocity averages 2.0 to
3.0 km/s in unconsolidated soft sediments, and 4.0–
5.3 km/s in the consolidated (hard) layer. A comparison
with a ‘site-specific’ (continental USA) vertical-section,
from Kearey and Vine, 1996 is also shown in Figure
11.18c, for comparison with the continental ‘crustal
types’ model in Figure 11.18a. One-dimensional
crustal models of Vp, Vs and density to 40 km depth for
the whole globe, continental crust and shelf, and the
oceanic crust are given in Figure 11.19.
Interesting insights into the local nature of crustal
reflections, and of course excellent velocity-depth data
from sonic logging and VSP have been obtained from
the KTB deep drilling project in Germany, where results
of 9.1 km of borehole logging and core analysis were
available in Harjes et al., 1997. The thirteen authors of
Figure 11.15 Average laboratory velocities for each rock type, at this paper related some interesting experiences about the
20 km equivalent depth, and at a temperature of nature of the strongest reflectors, which tended not to be
309°C. Christensen and Mooney, 1995. lithological boundaries alone, but fluid-filled fractures
and cataclastic fracture zones, sometimes associated with
such boundaries.
be commended for their extraordinarily far-reaching The most pronounced and discrete reflections were
review, from which we have borrowed many figures reportedly found in the compositionally homogeneous
in this chapter on crustal velocities. amphibolite unit, and originated from hydraulic fracture
A new global model for the earth’s crust, based on seis- zones at 4.0 and 4.8 km. Other reflections correlated with
mic refraction data published in the period 1948–1995 fluid-filled fracture zones in gneiss-amphibolite contrasts,
was provided by Mooney et al., 1998. The model was so uniqueness could not be determined. One may specu-
based on 5°  5° tiles (that measure 550  550 km at late that the hydraulic fracture zones had become propped
the equator). In each tile, crustal properties were in some way, perhaps due to sheared-dilated sections
described by seven layers: 1) ice, 2) crater, 3) soft sedi- (non-parallel to maximum stress), as discussed by Barton,
ments, 4) hard sediments, 5) crystalline upper crust, 1986, and extensively reviewed in Chapter 16.
6) middle crust, 7) lower crust. The predominantly gneiss and amphibolite sequences
The source location (mid-profile) of the numerous shown in Figure 11.20 showed lower Vp/Vs ratios in the
seismic refraction profiles for this monumental (2000 gneiss than in the amphibolite (due mainly to quartz
cases) study by Mooney et al., 1998 and others, are given content differences). In general, decreases of the Vp/Vs
in the world map in Figure 11.17. The triangles refer to ratios were caused by decrease of Vp rather than by an
the locations within continents, and on margins, where increase of Vs, which the authors liken to typical behav-
a velocity-depth function could be extracted from a iour in fractured, or jointed, and porous rocks.
published interpretation. The mid-point of a major The mean trends and individual results of sonic and
profile corresponds to the triangle location. In about VSP measurements down this unusually deep borehole,
10% of the cases the shear velocity-depth profile was are shown to at least 8.5 km depth in Figure 11.20. The
also reported. strong Vp (and Vs) velocity-depth gradient shows the
250 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.16 Average continental crustal velocities (shaded) compared to average laboratory velocities at simulated depths and temperatures.
Christensen and Mooney, 1995.

classic knee-shape, with average Vp increasing from (or Q c values) as low as 1, which signifies a good deal of
about 3 km/s close to the surface to about 5.3 km/s at jointing. The calliper log measurements also shown in
500 m depth and to about 5.7 km/s at 1000 m (based Figure 11.20a indeed support the idea of borehole walls
on mean VSP data). Comparison with the velocity- with a good deal of joint-related ‘overbreak’, which nor-
depth-Q (rock mass quality Q or Q c) curves shown in mally occurs when there are 2 or more joint sets (i.e. Jn
Figures 5.36 and 5.37, suggests that this upper 1 km of probably in excess of 4 or 6 in the Q-system of rock
paragneisses (with some amphibolite) may have Q-values mass quality description, combined with unfavourable
Velocity structure of the earth’s crust 251

Figure 11.17 Location map for the Mooney et al., 1998 global crustal model seismic refraction profiles. The triangles refer to the locations
within continents, and on margins, where a velocity-depth function was extracted from a published interpretation. In about
10% of the cases the shear velocity-depth profile was also reported.

(a) (b) (c)

Figure 11.18 Vp-depth structures for primary crustal types. a) Continental. b) Oceanic. Mooney et al., 1998. c) Continental USA veloc-
ity-depth section. Kearey and Vine, 1996.

anisotropic stress and presumably water pressure see: Nevada Test Site in the western USA, one may note that
Appendix A for Q-parameter ratings). nuclear ‘events’ of 155 to 1300 kilotons equivalent yield
Moving to an entirely different geology and location, were used for forward modelling of surface velocity data
and into an artificial ‘seismic’ environment, namely the that was recorded within 15 km of the underground
252 Rock quality, seismic velocity, attenuation and anisotropy

nuclear test explosions. Barker et al., 1991, used a


plane-layered structural model of the porous, low-den-
sity volcanic sequences beneath the Pahute Mesa to
derive the velocity-depth structures shown in Figure
11.21. These velocity structure models were needed to
determine the effective source functions of the under-
ground explosions.
The upper part of these velocity-depth trends show
broad similarity to the Vp-depth-Q c (normalised rock
mass quality) trends shown in Figure 5.37. Clearly the
porous, jointed volcanics are at the lower end of the Q c
range, or alternatively represent a ‘porous version’ of the Figure 11.19 Depth-velocity-density profiles from a crustal model
trends shown by Q c  1 to 10 in Figure 5.37. CRUST 5.1. Mooney et al., 1998. The predomi-
Before leaving continental velocity structure, we may nance of oceans causes the average and oceanic crust
look at two near-surface extremes, namely sea ice, or velocities to be low in the upper 3 km.

Figure 11.20 Borehole measurements and geological profile of the KTB super-deep well. Note (a) shows calliper log measurements and hole
diameter. Note the ubiquitous nature of faulting at all depths. Harjes et al., 1997.
Velocity structure of the earth’s crust 253

glacial ice and beach sand. These occur just above sea level While thin, floating sea ice, typically 10 m thick
and they have two aspects in common. They each display constitutes an approximately constant velocity layer, an
high gradients of velocity, but from different starting ice accumulation such as the Ross Ice Shelf, Antarctica
points. of many hundreds of metres thickness, displays a
sharply declining velocity gradient with depth.
A multi-layered upper 50 to 100 m called firn is
responsible for the steep velocity gradient. This includes
snow cover which becomes firn after one melt season, and
eventually becomes glacial ice when permeability to liq-
uid water drops to zero with subsequent burial.
Investigations using seismic reflection and seismic
refraction profiles, reported by Beaudoin et al., 1992,
were located on the 200 to 850 m thick Ross Ice Shelf
as shown in Figure 11.22.
The principal results of Vp versus depth are repro-
duced in Figure 11.23. Compressional wave velocities
in the near-surface ranged from 500 m/s at the surface
to 2000 m/s at 10 m depth, a gradient of 150 s1. From
10 m to 70 m depth, the velocity increased from
approximately 2.0 to 3.8 km/s, which represents a gra-
dient of 30 s1.
In this region, metamorphism of the firn is governed by
recrystallisation. Below about 70 m, any further com-
paction of the ice is by deformation of existing air pockets,
Figure 11.21 Velocity-density-depth trends for the Nevada Test
with little effect on velocity (though possibly giving an
Site Pahute Mesa. Barker et al., 1991. orthotropic distribution). Of the four compression wave

Figure 11.22 Location of Ross Ice Shelf seismic reflection and refraction profiles. Beaudoin et al., 1992b.
254 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.23 a) Cross-section of ice and water. b) Near-surface Vp-depth gradient caused by firn c) Overall Vp-depth profile, and chosen
models 1 to 4. Beaudoin et al., 1992b.

velocity models shown in Figure 11.23c, No. 1 was con- q  quartz density, air  air density,  is the porosity
sistent with the observed data. Below the ice, 570 m of of the mixture, assumed at the critical value of 40%.
water with a velocity of 1.44 km/s reached down to sub- Their estimates of theoretical velocity, and their low
sea sediments with a velocity of 2.7 km/s. measured values give velocities that are actually much
Conducting the shallowest possible high-resolution lower than the velocity of sound in air.
seismic reflection and refraction experiment in the upper
2 m of a sea-beach sand, Bachrach and Nur, 1998, meas-
ured a minimum P-wave velocity of 0.04 km/s. They 11.3 The continental margin
used only a 0.1 m distance between the shot and receiver. velocity structures
They calculated a theoretical minimum possible value of
0.013 km/s, considering the top few centimetres of dry Velocities at continental margins, such as that obtained
sand as a suspension of sand in air. in the Atlantic margin seismic experiment described by
The effective elastic modulus (Meff) and the velocity of Holbrook et al., 1994, naturally show some of the high-
the air-quartz mixture were calculated from the following est lateral variations of velocity, plus the familiar
equations: Vp  6.5 km/s to 10 or even 20 km depth beneath the
continental material. Figures 11.24 a and b show vel-
ocities and geological interpretation side by side, for a
1  1  (11.1)
  240 km section off the East coast of the US. Short black
Meff Mair Mquartz and white lines are reflectors.
The multi-channel data was acquired using a 177 litre
4 airgun array and 6 km long streamer, and coincident wide-
M  K    G (11.2) angle data from ten ocean bottom seismic instruments.
 3 
These seismic results along the US East Coast conti-
nental margin show the presence of a huge, high veloc-
eff  (1  )q    air (11.3) ity (7.2–7.3 km/s) igneous body of as much as
2.7  106 km3 in volume. This East Coast Margin
Meff Igneous Province (ECMIP) probably extends seaward,
Vp  (11.4)
eff making it one of the worlds really large igneous
provinces. The high velocity (in relation to thickness) is
nicely demonstrated in the two further versions of the
The following parameters were used: bulk modulus velocity-depth-distance sections, from Kelemen and
Kquartz  36.6 GPa, shear modulus Gquartz  45 GPa, Holbrook, 1995, shown in Figure 11.25.
Velocity structure of the earth’s crust 255

Figure 11.24 US East Coast continental margin velocities (a), densities in kg/m3 (b), and geology (c). Holbrook et al., 1994. Note that black
and white lines are reflectors.
256 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.25 Location and velocity-depth trends of two sections (BA-6, and EDGE-801) through the US East Coast margin, showing the
velocity-thickness anomaly. Kelemen and Holbrook, 1995.

11.3.1 Explaining a velocity anomaly This equation was subsequently corrected to lower-
crustal temperatures (400°C) using an assumed dVp/dT
In an effort to understand the likely composition of the gradient of () 0.0005 km/s/°C, by subtracting 0.2 km/s.
rock in this huge magnetic and seismic velocity anomaly, Figure 11.26b shows calculated Vp for rocks crys-
Kelemen and Holbrook, 1995, assembled numerous tallised from mantle melts as a function of the pressure
high pressure laboratory Vp data to try to differentiate of partial melting in the mantle. This was estimated by
the 25 km thick high velocity crust from the general Kelemen and Holbrook, 1995, using their relation:
8 km thick (Vp  6.9 km/s) sub-ocean crust.
Figure 11.26a shows a multiple linear regression fit Vp  6.712  0.16 Pmelting(GPa)  0.661 Fmelting
to 188 garnet-free, igneous and metamorphic rocks. (11.6)
Measured Vp at 25°C and confining pressures from 0.6
to 1.0 GPa, are compared with the bulk composition by where Vp  km/s and Fmelting is the melt fraction of the
weight of SiO2 or MgO in the samples. The empirical parental melt, using reported SiO2 and MgO contents,
relation obtained was: and the temperature corrected (0.2 km/s) version of
equation (11.5).
Vp  8.054  0.024 (%SiO2)  0.029 (%MgO) According to Kelemen and Holbrook, 1995, the
(11.5) goodness of fit of equation 11.6 did not substantially
improve when other oxides like FeO, CaO, Al2O3,
(where Vp is km/s). Na2O etc. were entered, because these compositional
Velocity structure of the earth’s crust 257

rates than lithospheric spreading rates, to produce the


necessary high pressure conditions.
Gravity anomalies at the surface may be an expres-
sion of non-hydrostatic stresses at depth, implying that
significant deviatoric stresses may exist. In the case of
the Hawaiian Islands, gravity anomalies associated with
flexure of the crust on either side of the Hawaiian ridge
are associated with average velocity reductions in Layer
2 of some 0.8 to 0.9 km/s within the flexural arch, some
155 km from the ridge (Brocher and ten Brink, 1987).
Elastic and elasto-plastic flexural models for the
region give predicted stress drops of 80 MPa in the upper
lithosphere. These authors compare this with a similar
confining pressure drop necessary to reduce velocities in
porous basalts by 0.5 km/s in the laboratory.
The lateral velocity variations to the north and south
of the Hawaiian ridge, produced partly as a result of
this flexure, are shown in Figure 11.27. The models go
from ocean floor to the bottom of Layer 2. We shall
see many more models of oceanic velocities in the next
section of this chapter.
Tomography was used by Hole et al., 2000, to invert
earthquake and air gun travel time data in the San
Francisco Bay area, to obtain 3D seismic velocity and
earthquake hypocentres. Most hypocentres were relo-
cated up to 2 km from their catalogue locations, and the
3D approach was also important for mapping lateral
velocity contrasts (subvertically through most of the
crust) where major strike-slip faults were present. These
lateral velocity variations correlated well with known
surface geology differences.
Strong velocity contrasts of 0.3 to 0.6 km/s were
observed in the middle crust when crossing the San
Andreas fault. Weaker contrast (0.1 to 0.3 km/s) existed
at other depths, and across two other faults. The relocated
seismicity hypocentres on the active strike-slip faults
Figure 11.26 a) Vp and mineral composition for 188 igneous and defined steeply dipping planes beneath the surface
metamorphic rocks at confining pressures of 0.6 to expression of each fault.
1.0 GPa. b) Empirical calculation of Vp versus pressure
Figure 11.28 compares a Hole et al., 2000, 3D based
of partial melting. Kelemen and Holbrook, 1995.
velocity tomogram with a 2D refraction model of
Holbrook et al., 1996.
Throughout the Hole et al., 2000, San Francisco Bay
variables were closely correlated with SiO2 and MgO in study area, no earthquake was found to occur in regions
the experimental set of data shown in Figure 11.26a. with Vp  6.3 km/s, and usually the various inversion
According to their analyses, the high velocity body tests produced a maximum velocity of 6.2 km/s. They
could have been produced from partial melting of man- surmised that the base of seismicity may be thermally
tle peridotite, using lower estimates of melt fractions controlled by a deeper brittle-ductile transition in the
(10%) but higher average pressures (2.0 GPa) than relevant Franciscan rocks. A simpler, shallow depth
that producing normal mid-ocean ridge basalt. They Q-M-Vp model argument (Barton, 1995, 2002) might
surmised active upwelling of the asthenosphere at faster be that fracturing, that has to be present, has kept Vp
258 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.27 Velocity-depth solutions from N and S of the Hawaiian ridge. Brocher and ten Brink, 1987.

below the 6.5 km/s ‘limit’ for intact, strong, highly cracks and joints, in other words, the joint sets that still
stressed rock masses. However there is of course a need remain partly open.
to extrapolate the Q-M-Vp model to greater depths to The author was able to recalculate the hypocentres of
be related even to shallow earthquakes. recent earthquakes using the improved three-dimen-
The amount of detail in depth-velocity structures for sional velocity models, which clearly differentiated the
onshore and offshore southern California (adjacent to the sedimentary basins from the nearby mountains. Hauksson
Pacific and North America plate boundary) was recently also compared his 3D Vp-depth profiles (using double
increased with new 3D Vp and Vp/Vs models using standard deviations) with laboratory Vp measurements
P and S-P travel times from local earthquakes and of triaxially loaded samples from McCaffree et al.,
from controlled sources (Hauksson, 2000). A 15 km 1998.
horizontal grid-spacing, and an average vertical grid It is of interest to note the ‘reluctance’ of the in situ
spacing of 4 km, down to 22 km gave new insight into data in Figure 11.31 to exceed Vp  6.5 km/s – which in
the heterogeneity of crustal structure in this earthquake- the Vp-Q-M model of Barton, 1995, is the supposed
prone region. The near-surface increase in P-wave veloc- limit for completely unjointed rock masses, or rock
ity, from the surface to 8 km depth was found to be masses with neither primary or secondary porosity and
rapid and had a logarithmic shape for stable blocks, but ‘normal’ composition (i.e. granites, gneisses etc.).
was slower and had a more linear slope for sedimentary Presumably the stress levels at 5 or more kilometres
basins (Figures 11.29a, b, c). depth are sufficient to completely close (in a joint-nor-
Ratios of Vp/Vs varied widely in the upper 5 km and mal direction) the apertures of any joints, since stresses of
often fell outside the typical ratio of 1.7 to 1.8 generally the order of 130 MPa and more are close to the expected
seen at lesser depths. Values as high as 1.9 to 2.0 were JCS values of joint walls in the schists, intrusives and
seen in sedimentary basins and in locations below an gneisses. (JCS  joint wall compressive strength, Barton
offshore channel (Santa Barbara). High Vp/Vs ratios and Choubey, 1977.) On the other hand with shearing
may be related to the high fluid content of near-surface deformation along the joints, apertures and permeability
Velocity structure of the earth’s crust 259

Figure 11.28 Comparison of a) 2D and b) 3D velocity tomograms for San Francisco Bay area crust. c) Range of velocity-depth models pro-
viding solutions to the inverted travel times. Hole et al., 2000.

could remain of finite magnitude at these (and greater) of 1, if we make no allowance for porosity n  1%
depths, as suggested by the work of Zoback and col- or c  100 MPa, or  100 MPa. The relatively low
leagues, reviewed in Chapter 16. velocities of some of the laboratory samples of schist
The 3D velocity data shows mean values of 5.4 and and intrusives (as low as 4.5 to 5.0 km/s at 1 km depth
5.7 km/s at one kilometre depth. In the Q-M-Vp model, equivalent) does suggest that n% and c adjustments
(see Figures 5.36 and 5.37), a significant amount of would be needed to fine-tune Qc rock quality estimates.
jointing and/or alteration along the joints would be There are other important details regarding the vel-
suggested, with a Qc (rock quality) value on either side ocity model for the San Gabriel ranges discussed above
260 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.29 Velocity-depth data for a) stable blocks, b) sedimen- Figure 11.30 Vp/Vs ratio trends for a) stable blocks, b) sediment-
tary basins and c) offshore regions off southern ary basins, and c) offshore regions of southern
California. Hauksson, 2000. California. Hauksson, 2000.
Velocity structure of the earth’s crust 261

11.4 The mid-Atlantic ridge velocity


structures

The object of systematic geophysical inversion techniques


is to derive structures which fit a given set of observations.
For many years, sub-oceanic marine seismic refraction
profiles were interpreted as a small number of layers
separated by planar interfaces, with a constant velocity
assumption for each layer. A major advantage of layer
solutions was that they could be developed using a desk
calculator.
As equipment (e.g. sonobuoys and repetitive sources
such as airguns) and computing capacity improved,
homogeneous layering assumptions from the 1960s,
e.g. Ewing, 1963, were first replaced by much finer
layering and then in the mid-1970s by continuous gra-
dients in velocity (e.g. Kennett and Orcutt, 1976). These
authors showed that the seismic data do not require
uniform layering as a solution, but they did not exclude
the possibility of homogeneous layering. A typical set of
their solutions, with bounds compared with the layered
solution is shown in Figures 11.32a, b.
The first geophysical downhole logging data for oceanic
crustal material is reportedly that of Kilpatrick, 1979.
He found that the predictions of low velocities from
refraction seismic were borne out by downhole sonic
logging. In situ sonic velocities were typically from 1.5
to 4.8 km/s in the upper 200 m of oceanic Layer 2A.
Calculated porosities of 13 to 41% were unexpectedly
high. Formation damage away from the drilled holes
was considered to be minor, as electrical resistivity away
from the hole showed a lack of radial variation. The
measurements were made in hole 396B (leg 46) of the
Deep Sea Drilling Project, near the mid-Atlantic ridge.
Figure 11.31 Comparison of velocity-depth trends for two onshore The reasons for the high porosities were interpreted as
regions of southern California (solid lines: ‘3D’) with being due to a combination of sediments, rubble, and
laboratory tests (all dotted lines). Hauksson, 2000 solid basalt in contrast to the compact nature of basalt
and McCaffree et al., 1998.
samples used in laboratory tests, which often has Vp
between 5.5 and 6 km/s and porosities from only about
(Figure 11.31b). Hauksson, 2000, warns that for shal- 2 to 8%. Open fractures and voids were assumed to exist
low earthquakes, the seismic waves from the hypocent- on a scale larger than the laboratory samples, giving high
res to the recording stations are travelling through the permeability throughout the drilled section.
schist with subhorizontal ray paths, while rays from A decade of Lamont-Doherty Geological Observatory
deep earthquakes may have steep angles of incidence. sonobuoy data led Houtz and Ewing, 1976, to conclude
Thus in both the Mojave (Figure 11.31a) and San Gabriel that the P-wave velocity of the sub-ocean crust at and
(Figure 11.31b) terrains, the rays from the hypocentres near ridge crests actually exhibited an increase in velocity
will sample the average velocity of the schist, because, with age. Numerous results from the Atlantic and Pacific
in relation to the foliation dip angles (45° to 90° for the shown in Figure 11.33 showed an obvious link between
Pelona schist) the rays have all possible azimuths and a Vp and age, up to some 40 million years. Deeper and
large range of take-off angles. older layers did not show systematic increase in velocity.
262 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.33 Measured velocities as a function of age for Atlantic


and Pacific sites. Houtz and Ewing, 1976.

The rocks concerned were generally pillow basalts or


vesicular, and could be weathered or massive and fresh.
The authors argued that it was difficult to envisage any
diagenetic change with the relatively low (effective)
overburden pressures, so filling of voids and cracks (and
presumably joint sets) with products of hydrothermal
mineralization must presumably be one of the mechan-
isms involved in the increasing velocity with age.
Different calculated porosities for 4 m.y. and 62 m.y.
crust as a function of depth, given by Whitmarsh, 1978,
and reproduced in Figure 11.34, suggested zero porosity
beneath 1.5 km of Layer 2A oceanic crust. The two rec-
tangles, the asterisk and the three dots were from limited
borehole data available at that time. Filling of cracks by
hydrothermal minerals with increasing age was cited as
the likely mechanism.
The early (and continued) difficulty of obtaining
samples of oceanic crust to several kilometres depth, led
Salisbury and Christensen, 1978, to ‘reconstruct’ the
intact rock Vp, Vs and dynamic Poisson’s ratio structure,
along a traverse through an on-land (Bay of Islands,
Newfoundland) ophiolite complex. The Vp and Vs data
shown in Figures 11.35 and 11.36 were derived from
hydrostatically confined and water saturated intact
samples. They suggested that the velocity structure should
be indistinguishable from normal oceanic crust, but with
the notable difference that the structure (voids, joints,
fractures etc.) were not of course sampled. The veloci-
ties, especially in the upper 1 km, therefore represented
maxima.
The authors showed from earlier studies (Christensen
and Salisbury, 1972), the strong link between velocity
and density for oceanic crust basalts. At high porosity,
with a density of only 2.5 g/cm3, Vp tended to be only
about 4.5 km/s, and from then on showed a linear
Figure 11.32 a) Velocity-depth bounds from inverted marine
refraction profiles, compared to Layer 2 and 3 con-
increase to about 6.5 km/s by the time the density had
stant Vp models. b) Mean value and error bars for reached 3.0 gm/cm3.
same profile CH-10A. c) Resolving kernels for CH- Clearly, seismic velocity and density are mutually
10A. Kennett and Orcutt, 1976. dependent properties of a rock, and each are strongly
Velocity structure of the earth’s crust 263

While density is a bulk property independent of


direction, seismic velocities can be anisotropic due to
the effects of microcrack alignment (e.g. in recovered,
stress relieved samples) or due to fabric anisotropy. We
may consider that density is related to the ‘hard poros-
ity’ of the rock, while ‘soft porosity’ in the form of
jointing (i.e. the rock quality Q-value) gives higher or
lower velocities depending on great depth or shallow
depth respectively.

11.4.1 A possible effective stress


discrepancy in early testing

Extensive laboratory testing of oceanic basement rocks


from deep drilling in the mid-Atlantic ridge by Hyndman
and Drury, 1976, highlighted the discrepancy between
laboratory seismic properties and in situ, bulk velocities
obtained from seismic refraction.
One of the problems was obviously sampling bias,
and the other may have been the early tendency to test
at much too high effective stresses. The ‘intact’ basalt
cores contain only disconnected vesicules and were
earlier assumed to be under a high effective stress due to
both sea load and crustal load. The actual effective stress
acting on deep-ocean shallow crust, is most likely
to be relatively low, due to a highly permeable assem-
Figure 11.34 Estimates of porosity-depth relations for different ages blage of basalt, sediment layers and joints/fractures and
of oceanic crust (basalt) given by Whitmarsh, 1978. larger voids.
The typical samples for laboratory tests were mini-
cores of 2.5 cm diameter drilled transversely to the
linked to porosity, uniaxial compressive strength and recovered pieces of 6 cm diameter vertical core. The
Young’s modulus. A major collection of density-Vp drillers operating onboard the Glomar Challenger drilling
measurements for a wide range of sedimentary, meta- ship generally recovered only 20% of the sections drilled
morphic and igneous rocks is shown in Figure 11.37, in mid-ocean upper-crustal rocks, so there was a further
from Ludwig et al., 1970, and P.J. Barton, 1986. The source of sampling bias.
studies with oceanic crust basalts cited above (Vp  Due to the controversial opinion of appropriate stress
4.5 km/s with   2.5 gm/cm3, Vp  6.5 km/s with levels for testing laboratory samples, discussed above,
  3.0 gm/cm3) clearly fit this huge data set. confining pressures of 17 MPa and even up to 207 MPa
The thin line in the centre of the scatter in Figure 11.37 were used by Hyndman and Drury, 1976, for on-board
is the mean velocity-density relationship, while the velocity measurements. A thin metal sheet jacketed the
heavy boundaries contain the great majority of data. At sample so the confining pressure was ‘close to the exter-
any given value, a density variation of about 0.2 to nal pressure, and the internal pore pressure was small.’
0.3 gm/cm3 and a velocity variation of 0.5 to 1.0 km/s Results of these high pressure velocity-density studies
are seen. The thick vertical bar, corresponding to a are reproduced in Figure 11.38.
density of 2.8 gm/cm3 is typical for crystalline conti- Hyndman and Drury, 1976, found that the vesic-
nental crust, with Vp  5.7 to 7.0 km/s. Closed circles ular porosity of their relatively fresh, young basalts,
represent sedimentary rocks and open circles represent caused Poisson’s ratio to increase with increased velocity
metamorphic and igneous rocks, relevant in these oceanic (i.e. from 0.28 at Vp  5.8 km/s, to about 0.3 at Vp 
crust studies. 6.4 km/s, due to the vesicules affecting Vp more than
264 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.35 Vp and Vs data from hydrostatically confined laboratory samples, plotted as a function of depth in the Blow-Me-Down mas-
sif of the Bay of Islands ophiolite complex Newfoundland. Salisbury and Christensen, 1978.

Figure 11.36 Smoothed envelopes of the same Vp and Vs test data, plus density and dynamic Poisson’s ratio, for the ophiolite Blow-Me-Down
massif samples. Salisbury and Christensen, 1978.

Vs. In contrast, Christensen and Salisbury, 1972, testing cause increased Poisson’s ratio with reducing velocities,
older and shallower depth basalts found Poison’s ratio in contrast to the decrease caused by equi-dimensional
reducing with increased velocity, due to the greater effect (vesicular) pore spaces.
of grain boundary weathering on Vs than Vp. The large Even in 100-m.y.-old sea floor the reduction in vel-
scale joints, fractures and voids in situ are also likely to ocity caused by weathering appeared to only extend to
Velocity structure of the earth’s crust 265

Figure 11.38 Vp-Vs-density data from high pressure triaxial tests


of three rock types. Hyndman and Drury, 1976.

the effective stresses are much lower than those assumed


or actually acting on the intact, but vesicular basalt.
Presumably the vesicules may originally be gas-filled
Figure 11.37 An extensive set of laboratory P-wave and density at a pressure at least as high as the surrounding water
data for sedimentary (closed circles), metamorphic pressure into which they were injected. Whether the
and igneous rocks (open circles). Ludwig et al., matrix porosity is too low to allow either inward satur-
1970, and P.J. Barton, 1986. ation with water (while under the ocean) or escape of
some of the excess gas pressure (when brought to atmos-
about 50 m depth. The mean laboratory velocity for pheric pressure) is perhaps still a point of controversy.
basalts younger than 20 m.y. and generally from deeper
than 50 m, appeared to be frequently in the range 5.9
to 6.0 km/s. Hyndman and Drury, 1976 showed labora- 11.4.2 Smoother depth velocity
tory velocities (of about 5.5 to 6.5 km/s) obtained with models
50 MPa confining pressure, next to the refraction-seismic
inferred velocity depth profiles. However, laboratory A new picture of the seismic structure of the oceanic crust
data for samples recovered from about 3 m to 60 m depth, began to emerge at the beginning of the 1980s, with the
given by Hyndman, 1979, gave velocities from as low work of Spudich and Orcutt, 1980a. It was found that
as 4 km/s to 5.2 km/s just below the ocean floor, up to velocity models in which velocity varied smoothly with
5.4–6.5 km/s at 60 m depth. depth generally explained wave amplitude variations bet-
These were also presumably tested at excessively high ter than the earlier ‘thick, homogeneously layered’ mod-
confining stress levels. The 2.5 to 3.5 km/s in situ vel- els. Some indications of this were apparent some 10 years
ocities were showing an apparent discrepancy of about earlier, but in this compilation, Spudich and Orcutt,
3 km/s relative to the intact rock, but in fact some of 1980a included many sites to confirm the new trends.
this difference was presumably due to the excessively The three homogeneous layers 2A, 2B and 2C sug-
high confining pressures applied to the intact samples. gested by Houtz and Ewing, 1976, were now con-
It is clear that most of the velocity discrepancy was due sidered too simplified, as finer structure, with significant
to the dominance of larger voids and fractures, and the lateral variations, showed a mix of velocity gradients, but
effect these had on the effective stresses. In a bulk sense, generally within the range 1 to 2 s1.
266 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.39a shows selections of the velocity depth


models assembled by Spudich and Orcutt, 1980a. The
upper diagram is for ten ridge or near-ridge sites, which
contrast significantly (at shallow depth) with the higher
velocities of the six older than 20 m.y. sites shown in
Figure 11.39b.
Spudich and Orcutt discussed possible reasons for the
quite steep velocity gradients (approximately 1 to 2 s1)
of oceanic ‘layer 2’ as being caused by finer structure,
which was highly variable laterally (i.e. it varied with
increased or decreased age). However the finer structure
was difficult to resolve with the currently existing reso-
lution of explosion seismology.
Drilling of the shallow crust from submersibles, that
had begun in the 1970s, had shown that the shallow crust
was permeated with numerous sediment and/or water
filled fissures, i.e. voids larger than laboratory sampling
scale. Spudich and Orcutt, 1980a argued that if ‘layer 2’
were, on average, composed of the Mid-Atlantic Ridge
basalt identified by Hyndman and Drury, 1976:

Vp  5.94 km/s

Vs  3.26 km/s

  2,80 Mg/m3

n  7.8%

then the addition of another 10% of porosity in the


form of larger water-filled cracks or fissures, could give
a Vp range from 5.5 to 2.6 km/s using current crack and
spherical pore models. This range could encompass nearly
the entire range of layer 2A and 2B velocities observed
by Houtz and Ewing, 1976.
The aspect ratio of cracks and fissures, whether they
were water filled or sediment filled, and whether they
could close in response to increased effective stress was,
naturally, the subject of much discussion. It was also
assumed by now that alteration of the older basalt could
have resulted in progressive infilling and cementation
of the cracks, thus explaining the increased velocities
with age.

11.4.3 Recognition of lower


Figure 11.39 a) Mid-ocean ridge and near-ridge (i.e. younger) effective stress levels
Vp-depth profiles derived by synthetic seismogram beneath the oceans
modelling. b) Vp-depth profiles derived from sites
older than 20 m.y. Spudich and Orcutt, 1980a. For On the subject of effective stresses in the uppermost
references to each profile, see their paper. permeable sub-ocean crust, Todd and Simmons, 1972,
Velocity structure of the earth’s crust 267

Table 11.3 Interpretation of upper ocean crust P-wave velocities


east of Guadalupe Island. Profile FFZ of Spudich and
Orcutt, 1980b. See Figure 11.40.

Layer Depth of crust (km) Vp km/s  (Mg/m3)

3 0.40 4.6 1.99


4 0.43 5.0 2.15
5 0.60 5.0 2.15
6 0.63 5.75 2.43
7 0.79 5.75 2.43
8 0.85 6.20 2.60
9 1.14 6.20 2.60
10 1.22 6.20 2.60
11 1.40 6.30 2.64
12 1.47 6.30 2.64
13 1.72 6.42 2.68
14 1.90 6.90 2.86
Figure 11.40 Best fitting Vp and Vs for profile FF2, described by
Spudich and Orcutt, 1980b.
and Spudich and Orcutt, 1980b, seem to have been
among the first to argue that it must have been the rap-
idly increasing effective stresses that were acting on the to 0.6 km/depth (i.e. 5.75 km/s). While discussing the
shallow sub-ocean crust that was causing the velocity likelihood of fractures and voids as well as matrix porosity,
increase. In other words, with Po as the water pressure at they focused more on matrix-type porosity changes than
the ocean floor (often 3 or 4 km depth of water), the pore on effective stress-induced closure of joints or fractures.
pressure Pp acting at depth z into the ocean crust will be:
11.4.4 Direct observation of sub-
Pp  Po  w g z (11.7) ocean floor velocities

while Pe the external stress at the same depth in the Three new experiments carried out on the Mid-Atlantic
crust will be: Ridge (MAR), near latitude 23°N, were described by
Purdy, 1987. The uppermost few hundred metres of
Pe  Po  r  g z (11.8) the oceanic crust were tested using a fixed ocean floor
hydrophone receiver, and a controllable explosive source
Therefore the effective stress Pp  Pe will be given by: that was towed within a few tens of metres of the rugged
bottom topography.
Pe  Pp  (r  w)g z (11.9) These 1 to 2 km refraction lines produced direct obser-
vation of the Vp structure of the upper 200 to 300 m
This obvious cause of an effective stress gradient in of the young igneous crust. One of the experiments
the crust was cited by these authors as a reason for was carried out over the site of Hole 648B of the Ocean
measuring velocities in laboratory tests down to zero Drilling Programme on a small volcano within the
effective stress, rather than the practice (at that time) of median valley of the MAR. This was close to a ‘zero age’
measuring velocity at elevated triaxial stress states with location, the two others were 14 km apart, above 7 m.y.
zero pore pressure. old crust. The latter gave higher layer 2A velocities than
Based on an interpretation of the sharp velocity gradi- the ‘zero age’ location.
ents shown in Table 11.3, (where a water depth of 3.4 km The sea floor velocity at ‘zero age’ was observed to be
has been subtracted) and based on Figure 11.40, Spudich 2.1 km/s, overlying an initial 4 s1, roughly 200 m deep,
and Orcutt, 1980b, argued that a vesicular (‘spherical’) linear velocity-depth gradient. The crust at this location
porosity of 18% would match velocities in the top 200 m consisted of fresh basalt lavas, with laboratory measured
of the basalt crust (i.e. 4.6 km/s) and that reductions of velocities in excess of 5.8 km/s (However these tests, car-
porosity to 2% could explain the increased velocity down ried out as early as 1978, were conducted at a confining
268 Rock quality, seismic velocity, attenuation and anisotropy

pressure as high as 50 MPa. Later tests in 1984 were also sets at 30° to 60° to the ridge axis of mid-Atlantic Ridge
at high 40 MPa confining pressures, suggesting about oceanic crust could explain their observed horizontal
6.0 km/s for the intact basalt, if ‘artificially’ confined at (azimuthal) velocity anisotropy of up to 0.4 km/s. There
too high pressure). was apparently a negligible seismic influence of perva-
On the basis of the above in situ/laboratory compari- sive ridge-parallel fractures on this anisotropy, which
son, Purdy, 1987, suggested that the 3.7 to 3.9 km/s the authors explained by their infilling by sediment, or
difference in velocities must be due to the presence of hydrothermal precipitation, or by their closure under
large-scale porosity, and various models were discussed. the high ridge-normal principal horizontal stresses.
It was inferred that from 30 to 50% porosity might be In the case of this mid-Atlantic crust of 1.1 to 3.4 m.y.,
needed using conventional arguments about pore shape White and Whitmarsh reported that the top of the
of that period. However, Purdy also referred to the basement had a velocity of approximately 3.7 km/s which
Todd and Simmons, 1972, and Spudich and Orcutt, increased on average at 1.0 to 1.2 s1 in the uppermost
1980b emphasis of the actual importance of effective 2.5 km, giving 6.0 km/s at about 2 km depth. The upper-
stress (subtraction of pore pressure from total stress), most 200–300 m showed higher gradients than this.
which had been used for a long time in soil mechanics, In relation to the Vp-Q-porosity-depth model (Figure
following Terzaghi’s theory of effective stress developed 5.36 in Part I), such velocities would suggest Q-values
earlier in the 20th century. Naturally, with exceptional of about 4 to 6 if the matrix porosity was about 5 to
water pressures of around 30 to 40 MPa, it is under- 10%. If we assume a mean ‘uniaxial’ compressive
standable that the theory of effective stress was appar- strength c of about 200 MPa, then a Q-value of about
ently late in being adopted in this hostile sub-ocean 2 to 3 is suggested, i.e. significantly jointed, perhaps
environment. with the following general character (see Appendix A):
Purdy considered the possibility (‘an elegant solution’)
that increasing differential [sic] (effective) stress could 50 2 0.66
be responsible for the 4 s1 velocity-depth gradient, Q    3 (11.10)
12 4 0.5
and presumed a 4 to 5% per 100 m porosity reduction,
as seen in the first 200 to 300 m of sub-sea layer 2A. In
fact experience from engineering (e.g. tunnelling) proj- At 1000 m depth a velocity of the order of 5.5 to
ects close to the surface, does suggest very high velocity 6.0 km/s would be predicted, if the rock at this depth
gradients when rock quality Q and Q c values (Barton, had unchanged character.
1995; 2002) are a) low and b) are rapidly increasing Christensen, 1984, investigated pore pressure effects
with depth, such as Q  0.1 followed by Q  1 and on basalts and dolerites and verified the strong effect of
very quickly Q  10 etc. i.e. nominal near-surface pore pressure variation on the velocities and on dynamic
‘jumps’ in Vp from 2.5 to 3.5 to 4.5 km/s with a super- Poisson’s ratios. The latter increased significantly as a
imposed stress increase effect on Vp as well. (See later result of increased pore pressure. He discussed the possi-
comments on the question of gradients of velocity, and bilities of over-pressured zones due to seals caused by
‘curve jumping’, i.e. increases of rock quality Q-values rapid accumulation of low permeability clays and shales,
with depth.) and also theorised that release of water accompanying
The reality below the near-surface zone is that both low grade metamorphic reactions in basalts could result
effective stress increases and clay gets compacted, and in excess pore pressure and resulting changes (reductions)
there is less clay as depth increases. Therefore one pro- in seismic velocities, and increases in Poisson’s ratios.
gresses rapidly from low rock quality Q or Q c to higher The authors noted that the pore pressure coefficient was
Q or Q c values quite quickly, with obvious conse- less than 1, and was not a constant for a given sample but
quences for increased Vp. depended on confining pressure and on pore pressure.
The differences between in situ velocity measurement
in the shallow oceanic crust and the higher matrix veloci-
ties measured at suitable (low) effective stress levels, is 11.4.5 Sub-ocean floor attenuation
obviously caused not only by moderate changes to the measurements
matrix porosity but by stress-sensitive (low aspect ratio)
jointing and fracturing. White and Whitmarsh, 1984, Reportedly the first direct measurements of Upper
found that sub-vertical, water-saturated conjugate joint Oceanic Crust compressional wave attenuation were
Velocity structure of the earth’s crust 269

Figure 11.41 Selected sub-ocean Qp profiles given by Jacobsen and Lewis, 1990. Variable attenuation, shows only partial consistency of
increasing Qp with depth.

described by Jacobsen and Lewis, 1990, using seafloor is an implicit relation between seismic Q and the rock
hydrophones and large (56 to 116 kg) explosive sources. mass quality Q-based deformation modulus Emass, or M,
The site was on 0.4 m.y. old crust, 13 km SE of the Juan when this is expressed in GPa. The above Vp of 5 km/s
de Fuca Ridge. At the same site a seafloor velocity of suggests a near-surface Q rock  32. Shallow, sub-ocean
2.7 km/s increased uniformly to 5.6 km/s at 679 m seismic Q of 10 to 20 might imply a significant degree of
depth, with gradients as high as 4.6 s1 at the surface ‘structure’, if equivalent Q rock values were, say less than 5.
and 4.1 s1 at depth. Values of seismic Q p obtained by Elsewhere, shallow ocean crust basalts have shown Q p
Jacobsen and Lewis, 1990, varied from 4 to 275, but values of between 20 and 50. Dry samples of oceanic
mostly clustered between about 10 and 20 in the upper basalts from layer 2, tested at (artificially elevated) con-
100 m, which was significantly lower than earlier esti- fining pressures of between 40 and 100 MPa have given
mates based on synthetic seismograms. Q p in the range 5 to 85. Differences are attributed to
They did not find a consistent increase of Q p with crack content, degree of alteration and matrix porosity.
depth, but several sets of data for 1/Q p did show such a These values are lower than the Q p values normally
trend of 1/Q p reducing with depth. (Figure 11.41) The obtained for sound basalts, where values of between 100
variations presumably might be connected with vari- and 600 can often be obtained (Wepfer and Christensen,
able degrees of fracturing or cooling joints, and partial 1990). Wepfer and Christensen, 1987, reporting the
closure with effective stress increases. Their results showed first laboratory measurements of Q p for dry and water-
that Q p was linearly related to frequency between 15 saturated oceanic basalts under appropriate pressure and
and 140 Hz, but frequency-independent components temperature conditions, showed Q p varying from 8 to
of attenuation were also evident. 100 at ultrasonic frequencies. The range was dependent
Pujol and Smithson, 1991, who analysed seismic wave on the state of alteration and porosity.
attenuation from VSP measurements in the Columbia The sudden steps up, and down from, very high
Plateau basalts, found values of Q p of about 50 (with in situ Q p values like 200–300, even negative 1/Qp
Vp  5.0 km/s) that were close to the value of Q p of 40 steps, leads one to question whether the early ship-board
found in Eastern North Sea basalt by Rutledge and triaxial test routines had an element of (local) correct-
Winkler, 1987. As has been argued in Chapter 10, there ness, meaning that some volumes of intact basalt can
270 Rock quality, seismic velocity, attenuation and anisotropy

perhaps be subject to high 30 MPa plus-rock-depth con- form of thin (  0.001) cracks significantly affect seis-
finement loads, interspersed by a majority of permeable mic velocities as these close, but this hardly affects over-
and low effective-stress-loaded permeable blocks. all porosities if thicker cracks and voids remain open.
Near the surface (depth A in Figure 11.42), where
both crack populations were assumed to be open, Shaw,
11.4.6 A question of porosities, 1994, estimated a Poisson’s ratio of 0.28. At intermediate
aspect ratios and sealing depth (B), he postulated that only the thicker (  0.1)
cracks and voids were open, resulting in an anomalously
Shaw, 1994, using Kuster and Toksöz, 1974 theory, pos- (and as observed) low Poisson’s ratio of 0.24. At greater
tulated that thin cracks preferentially close at shallow depth (C), all cracks were assumed to be sealed, return-
depth while lower aspect-ratio cracks do not. However, ing the velocities to that of the host rock, and Poisson’s
all crack populations were assumed to decrease with ratio was again about 0.28. In older crust, hydrothermal
depth. It was pointed out that even 0.1% porosity in the deposition caused thin cracks to seal first. Thicker cracks

Figure 11.42 Top: a) For young crust: thin and thick cracks at depth A; thin cracks are sealed at depth B, leaving only the thick cracks. b) For
old crust: all cracks are sealed. The above causes a Poisson’s ratio anomaly at depths of about 0.8 to 1.5 km, as shown in dia-
gram d) in relation to Vp and Vs data. Shaw, 1994.
Velocity structure of the earth’s crust 271

could remain unclosed and unsealed until the crust was value combinations’, and these may well be parameters
older, which then restored Poisson’s ratio to laboratory of convenience, as Al-Chalabi pointed out. In the rock
values. quality Q-system the ‘convenient’ parameters are clearly
those considered in the formulation of Q, rather than
additional parameters not thought of.
11.4.7 A velocity-depth discussion The non-uniqueness of the parameters in velocity-
depth functions, and the lack of physical significance of
The strong focus on velocity-depth data in these investi- any specific value of a given parameter had been over-
gations of the oceanic crust, in particular the supposedly looked up to then, according to Al-Chalabi, 1997.
‘anomalous’ velocities and gradients discovered in the An investigation of the velocity-depth gradients
mid-ocean fracture zones, should lead us to consider the that are synthesised in the Vp-Q-value-porosity-depth
fundamental non-uniqueness of velocity-depth relations, model of Barton, 1995 follows from Figure 5.36 (Part
as emphasised in a thought provoking article by Al- I), using a plotting format that can readily be compared
Chalabi, 1977. Figure 11.43a shows a smooth velocity- with the oceanic crust fracture zone data of Layer 2A
depth function such as: and 2B.
Figure 11.44 shows the results which were extracted
directly from Figure 5.36 for the case of six specific Q or
Vz  Vo  k.z (11.11)
Qc-values ranging from 0.001 (intensely fractured, thick
clay-bearing discontinuities) up to 100 (quite massive,
with actual small-scale fluctuations commonly seen in a unweathered competent rock mass with few widely
sonic downhole log. In this simple equation which is spaced joints, principally one set only). See Appendix A
attributed to Slotnick, 1936. Vo is the (P-wave) velocity for relevant – but non-unique – parameter ratings.
at the surface and Vz is the velocity at vertical depth z. The very steep Vp-depth gradients typically seen close
As pointed out by Al-Chalabi, 1977, the fluctuations to the ocean floor, in the first few hundreds of meters of
shown in Figure 11.43a, which represent actual vari- the new crust, could also be analysed with this near-surface
ability (and borehole effects) may not be seen in seismic based empirical method, developed mostly from civil
work, when the seismic wavelength is greater than the engineering and engineering geological projects. Note
scale of the fluctuations.
The actual variations of sonic velocity with depth can
be described by an extremely wide range of ‘parameter

Figure. 11.44 Vp-depth trends for six specific rock quality Qc


values, showing the assumed minor effect of poros-
ity (n) when Qc is high. Influence of n% (the ‘hard’
porosity) increases with lower Qc-values. Gradient
k  km/s/km  s1 is shown on the left-hand
side. Trends of Vp-depth-Qc were derived from
Figure 11.43 a) A linear velocity-depth fit to a unit showing fluc- Figure 5.36 (Barton, 1995). Note similarity with
tuations at sonic log scale. b) Surface velocity-gradient oceanic crust fracture zone data for Layers 2A and
trends for a unit logged at different depths in four 2B. The Qc-value represents mainly the ‘soft poros-
wells. Al-Chalabi, 1997. ity’, i.e. jointing.
272 Rock quality, seismic velocity, attenuation and anisotropy

the effect of porosities  1%, which increase strongly in C  mean gradient, 500 to 1000 m
influence as one moves from Qc  100 to the lower D  mean gradient, 25 to 1000 m
rock qualities (and larger ‘soft porosities’) towards the
left-hand-side of Figure 11.44. In the case of D, giving the overall gradient from V0
In practice there will be a tendency for increased to V1000, the separate effect of increased porosity is shown,
porosities close to the more weathered surface (arrow which moves curves successively to the right. In each
N trend), while with increasing depth, trend N will be of these four cases, a uniaxial compressive strength of
reversed and trend J may dominate (i.e. reduced joint 100 MPa (nominal) has been assumed (giving Qc  Q
frequency, mineral healing, increased Q-value and Qc in Figure 5.36). Higher values of c than 100 MPa, due
value, meaning that ‘curve-jumping-to-the-right’ will be to lack of weathering and low porosity would obviously
necessary). This empirically-based, near-surface method give higher surface velocities and a lower gradient
could perhaps help to explain ‘anomalously high’ gradi- k (s1), thereby plotting to the left of these four sets of
ents through Layers 2A and 2B. Both trend N and trend ‘100 MPa hard rock’ lines. The opposite would be the
J stimulate such an effect. case with c  100 MPa (i.e. with younger rocks).
Although ‘parameter value combinations’ in the rock
quality Q-system (Barton et al., 1974, Barton, 2002)
are definitely ‘non-unique’ (as per Al-Chalabi, 1997), 11.4.8 Fracture zones
a physically plausible situation is described by this
empirically-based choice of increasing ‘hard porosity’ The low velocity mid-Atlantic fracture zone studies
close to the surface, and reducing ‘soft porosity’ at depth reviewed in this section, show gradients of 3.0 to 3.5 s1
(higher Qc-values). for the upper 0.5 to 0.8 km, and seabed velocities as low
In a similar manner to the above, we can extract Vp- as 1.9 to 2.7 km/s. Reference to Figure 11.45 suggests
depth gradient (k) data from Figure 11.44, and express that curves B and D with suitably increased porosities
it in the simple form given by equation 11.11. The (nominal 1%) would fit such data very well.
results are shown in Figure 11.45. Data from mid-Atlantic Ridge fracture zone anom-
Four sets of data are shown in Figure 11.45: alies, discovered during the 1970s and 1980s, were
assembled by Detrick et al., 1993. (e.g. Figure 11.46)
A  mean gradient, 25 to 100 m (extreme) These emphasise the extreme heterogeneity of their
B  mean gradient, 100 to 500 m thickness and internal structure. In general, they consist

Figure 11.45 Analysis of Vp-depth gradients (k, s1) as a function of surface velocity Vo magnitudes, from the Vp-Q-value-depth-porosity
model, Figure 5.36, from Barton, 1995). Based on the Al-Chalabi, 1997 plotting format, shown in Figure 11.43b.
Velocity structure of the earth’s crust 273

of thin intensely fractured and hydrothermally-altered rise. Many of the names of researchers will be familiar
basaltic sections, overlying a rather shallow Moho. The after reading or perusing the last section.
sites of some of these investigations are shown in Figure Ewing and Purdy, 1982, both well known for their
11.46a, and a typical structural cross-section is shown in mid-Atlantic ridge studies, assumed a linear velocity-
Figure 11.46b. depth gradient in the upper 500 to 800 m of young (0
Velocity-depth trends for four of the large Atlantic to 4 m.y.) oceanic crust on the flanks of the East Pacific
fracture zones are shown in Figure 11.47. Initial vel- Rise. The data shown in Figure 11.48 indicates an aver-
ocity-depth gradients to 2 km depth appear to vary from age gradient of between 3.0 and 3.5 s1 for the upper
about 2 to 3 s1, though even steeper gradients are seen 0.5 to 0.8 km of oceanic crust, with seabed velocities
in the uppermost 100 to 200 metres. ranging from as little as 1.9 to 2.7 km/s. Ewing and
Purdy suggested that an even higher gradient might
exist in the upper 400 to 500 metres.
11.5 The East Pacific Rise velocity The evidence of very low velocities in the upper-most
structures oceanic crust was reportedly consistent with visual/
submersible and photographic evidence of pervasive
Following the forgoing summary of advances in under- fracturing in mid-ocean ridge crustal regions, where the
standing of mid-ocean ridges and fractured zones for basalt layer was exposed, and it was consistent with
the case of the mid-Atlantic ridge, we will now retrace drilling and logging results that showed high porosity
some of the steps made in studies of the East Pacific (Hyndman and Drury, 1976; Kirkpatrick, 1979).
Spudich and Orcutt, 1980a, had reasoned that a
10% porosity in the form of large fissures, added to a
measured matrix porosity of about 8%, could readily
produce a P-wave velocity of 2.6 km/s. For the case of
rubble zones, Hyndman and Drury, 1976, had esti-
mated a porosity of about 20%.
The Vp-depth data interpretation shown in Figure
11.48 indicated to Ewing and Purdy, 1982 that ‘the
percentage of cracks and voids’ diminished rapidly with
depth, giving a Vp of about 5.2 km/s at 800 m depth. If
we enter the Vp-Q-porosity-depth diagram shown in
Figure 11.44, at a velocity of 2.5 km/s, and at a matrix
porosity of 8%, we see a rock quality Qc value of about
1 (typical of weathered, heavily jointed rock).
At 800 m depth, with assumed unchanged rock mass
quality (but with higher effective stress), the P-wave
velocity is predicted to be 5.3 km/s, almost the same as
above, but without the linear-trend assumption. In other
words, the effect of increased depth may have largely
removed the porosity component created by the tectonic
and thermal fracturing and jointing, but need not have
removed (and indeed could not have removed) the
matrix porosity of a competent volcanic rock which
already had intruded into a pore pressure regime as high
as 30 MPa, resulting from a 3000 m ocean depth.
Ewing and Purdy, 1982 considered that their observed
data showed a significantly lower gradient of about 1 s1
Figure 11.46 a) Simplified tectonic map of North Atlantic fracture below 800 m, which would give a velocity of 6 km/s,
zones. b) Generalised velocity-depth structural cross- appropriate for the ‘solid unweathered basalt’, at about
section of a large Atlantic fracture zone. Detrick 1.5 km depth. They reckoned that this might be a
et al., 1993. reasonable maximum depth of significant fracturing
274 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.47 Velocity-depth data from four of the large Atlantic fracture zones. Shaded areas are for ‘normal’ oceanic crust from White, 1984.

although it could be argued that initial fracturing Q-Vp-depth-porosity-strength chart, in Figure 5.36,
extended to greater depth, and some healing has already taking the nominal 25 m line as ‘surface’, the measured
occurred there by compaction and/or cementation’. surface velocities of 3.05 and 2.4 km/s shown in Figure
Purdy, 1982, concurrently reported laterally homoge- 11.49 suggest rock quality Qc values of about 0.5 and
neous velocity-depth behaviour for two areas separated 0.08 (i.e. ‘very poor’ and ‘extremely poor’ engineering
by 110 km on the flanks of the East Pacific Rise. The tunnelling qualities), following which at 400 and 500 m
non-linear Vp-depth curves shown in Figure 11.49, depths, Figure 5.36 predicts velocities of about 4.7 and
which are averages for two areas of 20 to 40 km lateral 4.6 km/s, close to those measured. (Note that the empir-
extent, show, in this case, an inverse relation between ical Vp-Q-depth-porosity-strength relationship was
age and velocity, since the youngest crust has highest determined by trial and error, from land-based refrac-
velocity. Furthermore there was 100 m of sediment tion seismic in jointed and faulted rocks, and from both
overlying the older of the two sets of crustal data, which shallow and very deep cross-hole tomography measure-
would tend to add to the recorded velocity, yet it remains ments, each with Q-logging of relevant core).
lower than the youngest Vp-depth curve. The less steep gradients of about 2 s1 over the next
As a point of curiosity, the initial parts of the 500 m depth to 1 km were the result of measured vel-
curves to the ‘knees’ at 400 and 600 m have gradients ocities of about 5.8 and 5.4 km/s. These compare to
of about 4.2 and 4.4 s1. With reference to the predicted velocities of about 5.5 and 5.3 km/s, from
Velocity structure of the earth’s crust 275

Figure 11.49 Velocity-depth behaviour for 0.5 m.y.: crust (2E


Figure 11.48 Linear Vp – depth assumptions for the shallowest 2W), and 4 m.y.: crust (5E 6E 6W), from seismic
fi km of East Pacific Rize oceanic crust (ROSE area), refraction data on the flanks of the East Pacific Rise.
from Ewing and Purdy, 1982. They utilised a time- Purdy, 1982.
distance inversion method suggested by Dorman
and Jacobsen, 1981, which required linear Vp –
depth gradients in each layer. (OBH: ocean bottom
hydrophone).

Figure 5.36. Differences in porosity and uniaxial strength


between the youngest 0.5 m.y. and older 4 m.y. crust
could be used to further distinguish and interpret the
relative degrees of jointing, causing ‘curve jumping’
with increased depth, due to the likelihood of changed
rock quality Qc values with age.
Due to assumed aligned cracks, Shearer and Orcutt,
1986 found that travel times were affected by azimuth,
in measurements performed during the Ngendei exped-
ition to the South Pacific. They estimated 0.2 km/s
difference in P-wave velocities and 0.1 km/s difference
in S-wave velocities in the upper 1.5 km, caused by
azimuth. In the upper mantle, from about 7 km depth
below the sea bed (Figure 11.50), the difference in
P-wave velocity was about 0.45 km/s, but a nearly Figure 11.50 Anisotropic velocity versus depth model which satis-
isotropic S-wave was indicated. fies the Ngendei, South Pacific data. Solid line:
They interpreted the crustal anisotropy by a model NNE velocity, dashed line: ESE velocity. Shearer
involving aligned cracks parallel to the original spreading and Orcutt, 1986.
276 Rock quality, seismic velocity, attenuation and anisotropy

ridge, resulting in a fast direction perpendicular to the


fossil spreading direction. The upper mantle anisotropy
was consistent with there being aligned olivine crystals, in
which the fast direction was parallel to the fossil spreading
direction.
The problems posed by zero-age oceanic crust with
Vp  2 km/s, compared to about 6 km/s for intact basalt
continued to provide challenges for theoreticians and
practitioners working on the origin, formation and
structure of mid-oceanic crust. Studies resembling Mid-
Atlantic Ridge theories about hard and soft porosity
(low aspect ratios) and preferential mineral sealing, were
also performed with East Pacific Rise data.

Figure 11.51 Vp-depth structure for zero-age crust, 20 ka crust


11.5.1 More porosity and fracture (east and west) and 120 ka crust near the East Pacific
aspect ratio theories Rise axial summit graben. Christeson et al., 1991
and Purdy et al., 1991.
Low aspect ratio cracks, and their reduced frequency of
occurrence and reduction in aperture with depth, and
probable sealing with hydrothermal minerals in the
case of older oceanic crust, were some of the variables
confronting those researching the variable structure of
mid-ocean crusts.
Using theories termed extended-Walsh and extended-
Kuster-Toksöz, Berge et al., 1992, utilised a range of crack
aspect ratios ranging from extremes of 0.5 to 0.001, for
depth zones ranging from 0 to 500 m below the sea floor,
and succeeded in matching the Vp-depth trends for 0,
20 and 120 ka (1 ka  1000 years) oceanic crust from
Christeson et al., 1991, and Purdy et al., 1991. These are
shown in Figure 11.51. The method of Berge et al., 1992,
was one of data fitting, not forward prediction.
Berge et al., 1992 theorised that for 120 ka material
with Vp  2.5 km/s, porosity should lie between 24 and
34%. Slower (Vp  2.2 km/s) zero-age crust was less
well-bounded; a porosity of between 26 and 43% was
predicted. The extended Walsh model used by Berge
et al., 1992, required porosity-depth distributions for
the various crustal ages, as shown in Figure 11.52.
Wilkens et al. 1992 managed to match Ocean
Drilling Program/Deep Sea Drilling Project (ODP/
DSDP9 Hole 504B and Hole 418A Vp – depth data, to
500 m depth, by modelling cracks of small aspect ratio
that, in a ‘fast’ model became sealed if of sufficiently
small aspect ratio, and in a ‘slow’ model did not seal. Figure 11.52 Theoretical variation of porosity for matching
Deeper in the profile they ‘closed an increasing volume Vp-depth data, with extended Walsh crack aspect
of lower aspect ratio pores’. ratio fit. Berge et al., 1992.
Velocity structure of the earth’s crust 277

11.5.2 First sub-Pacific ocean core In general the uppermost 100 m was an aquifer of
with sonic logs and rubbly pillow basalts, breccias and a few massive flows,
permeability tests and greatest variability and largest velocity gradients
occurred here. The next 0.5 km was composed of pil-
At the beginning of the 1980s, in a sub-ocean Deep Sea low basalts, flows and breccias with an abundance of
Drilling Project borehole in the eastern equatorial Pacific minerals and alteration products. Basalt dikes were
ocean, in the Costa Rica Ridge area, it was possible for typical in the lower 350 m. Velocities, porosities and per-
the first time to correlate core (but usually of low % meabilities varied approximately as shown in Table 11.4.
recovery) with downhole sonic logs, borehole televiewer The fact that average recovery of core was only 20%
logs, and permeability test results. This was first per- suggests many vertical and sub-vertical discontinuities
formed to a depth of 1 km, through layers 2A, 2B and were not sampled. Several of the well logs suggested the
2C. A schematic section and downhole logging results presence of zones of intense fracturing and open porosity,
from Newmark et al., 1985, is shown in Figure 11.53. but the reducing permeabilities with depth clearly sup-
Based on the vertical borehole logging (i.e. biased ported the general observation of increased mineral sealing
against vertical structure) the upper 50 metres contained with depth, and presumably increasing effective stress
numerous horizontal to sub-horizontal fractures, thick effects as well. (Of course the second leads to the first, if
basalt flow units, and thin interbeds of pillow structures. finest fractures are preferentially sealed).

Figure 11.53 Downhole (504B) sonic velocities and schematic structure of 1 km of oceanic 2A, 2B and 2C crust, from the equatorial east-
ern Pacific (Costa Rica Ridge area). Newmark et al., 1985. Note sediment and rock velocity contrast.
278 Rock quality, seismic velocity, attenuation and anisotropy

Table 11.4 Approximate velocities, porosities and permeabilities from downhole measurements in the top 1 km of Hole
504B, equatorial Pacific ridge. After Newmark et al., 1985.

Zone Vp n% k (cm2) K (m/s) Vp -depth gradients

Upper 100 m 3.7 4–10 109–1010 106–107 Steepest


Middle 550 m 4.8 – 1011–1012 108–109 Moderate
Lower 350 m 5.6 1.5–4 1012–1013 109–1010 Moderate

permeability-depth measurements, e.g. to 1 km depth


in igneous and metamorphic rocks. (Barton, 2002).
The Troodos ophiolite estimates of van Everdingen,
1995, suggested a decrease in joint aperture and trace
length with depth, and very marked anisotropy. The
interpreted principal permeability directions in layer
2B oceanic crust given by van Everdingen, 1995, are
illustrated in Figure 11.54. These land-based measure-
ments appeared to have been at least partly based on the
apertures indicated by epidote, quartz and later calcite
fillings.
Based on the usual inequality of hydraulic aperture
Figure 11.54 Principal permeability directions in layer 2B oceanic (e) and the (rough-walled) average physical aperture (E)
crust, showing anisotropic principal permeability
(i.e. E  e, Barton et al., 1985), the above method of
magnitudes K1, K2 and K3. Interpretation based on
estimating apertures could explain the higher estimates
observations of fracturing and mineralization of an
(on land) ophiolite in Cyprus. van Everdingen, 1995.
of permeability (e.g. 103 m/s), which would also be
reflecting the negative effective stress episodes that would
necessarily occur during hydrothermal penetration of
fluids. Such was probably not typical of the effective
The in situ bulk permeability of extrusive volcanic stress conditions in operation when the above DSDP
rocks measured in the Deep Sea Drilling Project (DSDP) permeability measurements were made, since the per-
drillhole 504B (Costa Rica Rift area), was subsequently meabilities were of only moderate magnitude.
quoted as 4  1014 m2 (or about 4  107 m/s) at (If effective stresses were locally negative during the
20°C, which corresponds to that of a well-jointed rock DSDP permeability testing – a hazardous boundary
mass with a rock quality Q-value  0.25, based on a condition for drilling of wells – then the resulting larger
conversion of the above permeabilities to a Lugeon apertures, e.g. 1 mm or more, actually giving much
value of about 4 (see Chapter 9). In contrast, perme- higher permeabilities, would then have satisfied the
abilities of unfractured basalt, diabase and gabbro may assumed ‘mineral-filled opiolitic boundary condition’
lie in the range 1016 to 1023 m2 (or about 109 to of e  E.)
1016 m/s). A combination of deep sea crustal permeability meas-
Systematic investigations of jointing characteristics urements and interpretations of mineral-filled ‘frozen’
in ophiolitic (on land) remnants of oceanic lithosphere apertures from the Troodos ophiolite are shown in
were used by van Everdingen, 1995, to infer the possi- Figure 11.55.
ble joint structure effects on permeability in layer 2B of Similar, ophiolite observations of fossil flow porosities
the oceanic crust. A compilation of permeabilities meas- and permeabilities, based on cubic law calculations using
ured or inferred for sub-sediment pillow lavas (from mineral filling thicknesses (e.g. Norton and Knapp,
about 250 to 900 metres beneath the sea floor) and for 1977), were reported by Nehlig and Juteau, 1988. With
the underlying sheeted dike complex (from 900 apertures of 1, 2 and even 5 mm (e.g. of epidote), it is
to 1600 m) showed a range from 1010 to 1018 m2 (or clear that artificially high estimates of ‘permeability’ are
103 to 1011 m/s). This happens to be comparable to made, such as many estimates of 1010 m2, or about
the usual maximum range of measured land-based 103 m/s. These are exceptionally high. If many such
Velocity structure of the earth’s crust 279

Figure 11.55 An interpretation of possible permeability trends in the first 1600 m of ocean crust, based on parallel-plate modelling, with
matrix addition, plus measured permeabilities from various sources. See van Everdingen, 1995, for references.

apertures were caused during negative effective normal 11.5.3 Attenuation and seismic Q due
stress episodes, as seems likely in sill and dike intru- to fracturing and alteration
sions, they would not reflect ‘virgin’ permeabilities, as
existing prior to the hydrothermal fluid injections. On Swift et al., 1998a described the seismic attenuation, for
the basis of this reasoning, the ‘fossil’ apertures observed the upper 1.8 km of Hole 504B (Costa Rica Ridge area:
in recovered core may not accurately reflect the porosity see upper 1 km of permeability data in Table 11.4 from
available at the time of hydrothermal fluid injection. Newmark et al., 1985). About 60% of the total observed
280 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.56 VSP and sonic log measurements in upper 1.6 km of Hole 504B. VSP Poisson’s ratio log also shown. Qi (intrinsic) values are from
Swift et al., 1998a with Qs (scattering) values at 10 m intervals from Goldberg and Yin, 1994. Note the very high dynamic
Poisson’s ratios at shallow depth, derived from the VSP Vp and Vs values, due to thin flows, pillow lava, and breccia.

amplitude decay apparently occurred in the pillow basalt, ● 1 Ma Qseis  20 to 60


due to geometrical spreading and impedance contrasts, ● Near ridge axis Qseis  10 to 20 (near sea floor)
and much of the remaining amplitude decrease was con- ● Within first 0.4 Ma Qseis  50 to 60 (near sea floor)
centrated in two layers, about 500–650 m and 800–900 m ● few Ma Qseis  300 (1 to 2 km depth)
below the ocean floor, as shown in Figure 11.56. ● Lab samples of basalt at 50 MPa confinement:
Attenuation in these layers was described by low seis- Qseis  55 to 120
mic Q of 10 and 8 respectively, due to intrinsic attenu- ● Lab samples of diabase dikes at 100 MPa confine-
ation mechanisms. In the case of the upper zone the ment: Qseis  70 to 370.
authors believed that alteration-mineralogy may have
been responsible for the attenuation, as there was no Other attenuation data from the upper crust in these
crack-related reason for the high attenuation. In the mid-ocean ridge structures, show seismic Qp estimates
past this may have been a zone of high porosity prior to varying from 35 to 80 in general (Harding et al., 1989;
pore-space filling by zeolites. Vera et al., 1990; Wilcock et al., 1992) and dropping as
In the case of the lower zone with seismic Q of only low as 11 to 20 in the uppermost crust (Christeson et al.,
8, the authors mention that this 800–900 mbsf interval 1994). Kappus et al., 1995 found that their reflectivity
coincided with features in core descriptions and logs synthesis computed with seismic Qp varying from 50 to
that suggested the presence of an intracrustal deforma- 100, fitted original data very well.
tion zone, or a sub-horizontal fault. Tectonically dis- The high attenuation, low seismic Q zone described
turbed rock, a local minimum in resistivity, a maximum by Swift et al., 1998a, had sonic log velocities down to
in inferred porosity, a decrease in Young’s modulus, 1 km/s lower in this fractured interval. The Qi intrinsic
increased fracturing in the wall of the borehole, low attenuation and Qs scattering attenuation interpretation,
aspect ratio cracks containing fluid, are among the various the corresponding sonic and resistivity lows and the
descriptions of this low seismic Q zone. It could have dynamic Poisson’s ratio determined from VSP, are each
been helpful to also know the rock mass quality Q, reproduced in Figure 11.56.
or at least RQD, among all these mostly qualitative Swift et al., 1998b also referred to the large-scale resis-
descriptions. tivity measurements at Hole 504B, giving parallel esti-
Attenuation studies in upper ocean crust cited by the mates of bulk porosity reduction with depth.
authors, indicated a certain age relationship, with atten- The increase in Poisson’s ratio is caused by dispropor-
uation greatest and seismic Q lowest in the youngest rock, tional reduction in S-wave velocity compared to P-wave
as reviewed by Swift et al., 1998a. velocity, which theoretical studies by Shearer, 1988, have
Velocity structure of the earth’s crust 281

shown should occur with relatively thin cracks or joints, about 800m depth, failed to detect the low velocity that
having aspect ratios less than 0,005 in an otherwise would ‘normally’, (nearer the surface), be an obvious
isotropic solid. feature of such a fault: see Figures 8.25 and 8.26 in Part I.
Swift et al., 1998b gave an interesting comparison of
laboratory data (open circles in Figure 11.57) obtained
apparently from 100 MPa confinement, with Vp, Vs and 11.5.4 Seismic attenuation
dyn trends obtained from VSP. The 100 MPa tomography across the East
confinement applied in earlier laboratory studies was an Pacific Rise
incorrect simulation of an actual much lower effective
stress gradient, as discussed earlier in this chapter. Possibly Some results of the first three-dimensional tomographic
some effect of a correct effective stress gradient might have study of crustal seismic attenuation across the East Pacific
been observed on the intact samples, if this had been Rise near 9°30N, dating from 1988, was reported by
applied. We can see that, in relation to this presumably Wilcock et al., 1995. The rather unique layout of 480
unrealistic 100 MPa confinement, there is up to 2 km/s explosive charges distributed over an ocean-bottom
deviation (reduction) of in situ P-wave velocities due to 16  16 km grid is reproduced in Figure 11.58. Solid
structure, at the top of the hole. The difference is more symbols are ocean bottom receivers, which included
than 3 km/s in the case of S-waves, presumably due to the analogue and digital hydrophones and seismometers.
water-filled structure close to the ocean floor. Bathymetric contours (m) are also shown. The East
Swift et al., 1998b suggested that large-scale, well- Pacific Rise near 9°30N is a fast spreading ridge, charac-
oriented vertical fractures (i.e. a joint set) formed tectoni- terized by a sharp upper-crustal to mid-crustal velocity
cally, did not have a detectable effect on P-wave velocities. inversion some 1.5 to 2 km below the seafloor, presumed
Presumably this is an expression of the effect of tight to be the roof of an axial magma lense. Small mid-crustal
closure by stress, which has also been observed in the case (i.e. 3 to 5 km deep) magma chambers appear to be a
of an obvious fault ahead of a (stuck) TBM tunnelling common feature of these fast-spreading ridges. Since a
machine. (Seismic velocity tomography performed from narrow lense of partly melted rock would solidify rapidly
diverging holes ahead of the particular tunnel face at in a cooling environment dominated by hydrothermal

Figure 11.57 Comparison of VSP (in situ) velocity structures


in Hole 504B on the Costa Rica Rift, with numer-
ous researcher’s laboratory data (open circles), Figure 11.58 A 16  16 km grid with 480 explosive charges (open
which show little effect of depth with the (artificial) circles) and the ocean bottom receiver array (sold
applied effective stress of 100 MPa. Swift et al., symbols). Bathymetric contours (m) are also shown.
1998a. Wilcock et al., 1995.
282 Rock quality, seismic velocity, attenuation and anisotropy

stress enhancement above the lense, because 20 km off


axis, seismic Q remains at only 45, despite greater age at
increasing distance from the ridge.
Inversions for individual receivers showed that seismic
Q increased from average values of 40–50 in the upper
1 km, to at least 500–1000 at depth greater than 2 to 3 km.
Results appear to be in agreement with other studies of
attenuation in young oceanic crust. In an on-bottom
refraction study near the centre of this tomography
experiment, Christeson et al., 1994, measured seismic
Q in the upper layer 2A of only 11 to 22.
Wilcock et al., 1995, cited three dominant attenuation
mechanisms in stable continental crust and marine sedi-
ments as: intrinsic attenuation as a result of Coulomb fric-
tion along cracks (Walsh, 1966), the flow of pore water
(Biot, 1962) and scattering. They suggest that these mech-
anisms, together with scattering from the rough seafloor,
may also be important in young, igneous oceanic crust.
Laboratory measurements on oceanic basalts (Wepfer and
Christensen, 1990) suggested that attenuation would also
increase with porosity, degree of alteration and water
saturation.
One may append here the more specific and pre-
sumed ‘micro-shearing ‘and micro-flow’ terms to qualify
the above ‘friction’ and ‘pore water flow’, since there are
physicists who visualise only sub-atomic magnitudes of
deformation across microcracks with the passage of seis-
mic waves, and others who even deny that friction can
be involved in attenuation.
The subject of attenuation is indeed controversial,
but there are clear indications, presented in Chapter 10,
and further discussed in Chapters 13, 15 and 16, that
Figure 11.59 Results of the inversion for a vertical cross-section
micro-shear and micro-normal deformations along/
and an along-axis slice, showing the lower crustal,
across attenuating cracks and joints (i.e. displacement
sub-ocean ridge, Qp1 attenuation structure. The
along-axis result is at 4 km depth. The Qp values
discontinuities), as interpreted from seismic anisotropy
range from 25 to 100. Wilcock et al., 1995. field measurements, are closely following the stiffness
(or compliance) magnitudes seen in the ‘static’ macro-
deformation (stress-closure and shear-displacement-
circulation, it is generally considered that the supply of dilation) testing of joints that is more common in rock
magma from below must be relatively steady-state and mechanics. Dynamic compliances are often smaller
uniform, deep beneath the axis of the ridge. than the inverse of ‘static’ stiffnesses, but only margin-
The attenuation expressed as Q1, (or Q1 p ) for a cen- ally so in the case of normal stiffness in rock masses of
tral vertical section, and for a horizontal cross-section at good quality. They seem to be of the same order of
4 km depth is shown in Figure 11.59. The four attenu- magnitude, or even closer. Dynamic shear stiffnesses
ation (Q1) values give estimates of seismic Q  25, may be up to two orders of magnitude stiffer than static
33, 50 and 100 from the central lense (with an assumed shear stiffnesses. This seems hardly enough difference
few percent of melt) to 2 and 3 km off-ridge distances. to prevent friction from being involved in attenuation.
In the upper 1 km by contrast, seismic Q averages Since fractures are preferentially oriented parallel
about 35 off-axis, increasing to 65 near the axis rise. One to mid-ocean rise axes, it has been postulated that
may speculate that this might be due to a horizontal attenuation should be higher for waves propagating
Velocity structure of the earth’s crust 283

perpendicular to the rise (Macdonald, 1982). fault and over 10 km from the nearest second order
Unfortunately there were insufficient axis-parallel paths discontinuity.
in the tomographic investigation of Wilcock et al., The main features revealed were a thin low velocity
1995, to distinguish the two orientations. layer 2A consisting of about 80 m of (nearly) constant
velocity rock (2.45  3% km/s at the sea floor) fol-
lowed by a steep gradient through 150 m of rock to the
11.5.5 Continuous sub-ocean floor base of layer 2A. The thickness of the 4 km/s iso-velocity
seismic profiles contour was mostly 130  20 m, increasing to 180 m
towards the north. This implied a maximum Vp-depth
As time went by, investigations of the mid-oceanic gradient of about (4.0  2.45)/0.05  31 through a
ridge areas become even more extensive with the added 50 m section compares with the also very steep gradients
possibility to compare new results with ever more numer- at shallow depth in jointed (sub-continental) rock masses
ous earlier studies. An integrating report of this nature by shown in Figure 11.44, as derived from the Q-Vp-
Kappus et al., 1995, also described a high-resolution seis- depth-porosity-strength model of Barton, 1995, 2002).
mic velocity profile of the uppermost 500 m of East Figure 11.60 shows the velocity of the top layer
Pacific Rise crust at 13°N, along a 52 km segment of the 2A (mean 2.45 km/s) and the thickness of the 4 km/s
ridge crest. The continuous profile, synthesised from 70 iso-velocity contour (mostly 130  20 m). Velocities at
individual 1-D models spaced at 750 m, showed remark- the top of layer 2A and at the top of layer 2B are shown
able lack of variation. The 53 km segment was however in Figure 11.61. A reflection deeper in the crust (tri-
more than 100 km from the nearest first order transform angles) at a velocity of 5.5 to 6.1 km/s is also shown.

Figure 11.60 a) Velocity at top of layer 2A (mean 2.45 km/s) and b) thickness of layer 2A where the 4 km/s iso-velocity contour is found (130
to 180 m). Kappus et al., 1995.

Figure 11.61 Velocities at top of layer 2A (circles) and at top of layer 2B (squares). A reflection deeper in the crust (triangles) was interpreted
as the lid of a magma chamber. Kappus et al., 1995.
284 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.62 Seismic velocity structure of the upper 500 m of a 52 km long segment along the mid-oceanic ridge crest of the East Pacific
Rise. Kappus et al., 1995.

(a) (b)

Figure. 11.63 Average P-wave velocity-depth curve and standard deviation for 70 profiles covering a 52 km segment of the ridge crest, to a
depth of 400 m. Kappus et al., 1995.

The relative uniformity of much of the 52 km long the seafloor. This transition to an entirely different (2B)
segment (measured at 70 locations) is further emphasised gradient is seen more easily in Figure 11.63, which
by the strikingly beautiful contoured velocities shown shows the average velocity depth behaviour (solid lines)
in Figure 11.62, which show rapid increases in velocity and the mean of N and S parts of the segment (dashed
at first followed by slower increases due to longer depth lines). A starting model of velocity versus depth, and
intervals. various iterations is shown in Figure 11.64.
The base of layer 2A was assumed to be the lower For the purpose of estimating gradients, the average
part of the steep velocity gradient at about 230 m below velocity-depth data is reproduced in Table 11.5. This data
Velocity structure of the earth’s crust 285

Table 11.5 Average velocity-depth data from 70 profiles along a


52 km segment of the East Pacific Rise ridge crest.
Kappus et al., 1995.

Depth below P-wave velocity Standard deviation


seafloor, m km/s km/s

0 2.25 0.000
40 2.38 0.135
80 2.57 0.240
120 3.17 0.350
160 3.82 0.323
200 4.36 0.289
240 4.71 0.217
280 4.88 0.158
320 5.00 0.140
360 5.10 0.140
400 5.19 0.138
440 5.26 0.141
480 5.36 0.172

estimates varying from 35 to 80 in general (Harding


Figure. 11.64 One of the starting models for Vp-depth inversion, et al., 1989; Vera et al., 1990; Wilcock et al., 1992).
to 1 km depth below the ocean floor. Successive iter- Kappus et al., 1995 found that their reflectivity synthe-
ations are shown, the fourth with asterisks. Kappus sis computed with seismic Qp varying from 50 to 100,
et al., 1995. fit their original data very well.
The low velocities of seafloor, age zero, mid-oceanic
has been plotted among the ‘soft porosity’ (joint-related) ridge crest materials from numerous studies in the
curves of the Qc-Vp-depth-porosity-strength model, period 1976 to 1994, reviewed by Kappus et al., 1995,
reproduced for easier comparison in Figure 11.65. had the following values in km/s: 2.5, 3.1, 2.1, 3.5, 2.35,
The Table 11.5 data plotted in Figure 11.65, shows 2.2, 2.0, 2.45, 2.4 and 2.7 (read from the zero-age end
strong evidence of structural (and matrix porosity) effects, of eventual ranges of velocities). The mean value of
and much rock quality improvement with depth increase 2.5 km/s implies a near-sea floor rock quality Q-value
(i.e. curve-jumping), in the upper 250 m. Interestingly of only 0.1, as also roughly indicated in Figure 11.65.
the Vp data below this depth suggests typical jointed rock Velocity-depth gradients for layer 2A as a whole appear
Q-values (Barton et al., 1974; Barton 2002) in the range to have ranged from 3.5 to 5.5 s1, though they do not
2 to 8 down to 1000 m depth, with data paralleling the appear to have been quoted in many of the papers ref-
trends of Q  1 and Q  10 suggesting effective stress- erenced by Kappus et al., 1995.
joint-closure effects. A uniaxial strength of 100 MPa These low velocity, zero-age, crustal values have to
has been assumed in this example. If the basalt was closer be contrasted to laboratory velocities for young basalts
to 200 MPa strength, a uniform shift to half as high of at least 5.6  0.4 km/s (e.g. Hyndman, 1976).
Q-value (i.e. from about 5 to 2.5) would be involved. Information from drill holes (Alt et al., 1986; Nehlig
A plausible Q-parameter construction to explain such and Juteau, 1988) reinforce the idea that the low veloci-
values of rock quality beyond 250 m depth could be the ties are strongly linked to structure (i.e. discontinuities,
following: joints, fractures) and to matrix porosity, since there is
evidence of strong circulation of hydrothermal fluids,
60 3 0.5 mixing of cold and hot fluids and alteration, which could
   3.8
12 4 0.5 be intense at some levels.
Collier and Singh, 1998, utilised wide-aperture seis-
Attenuation in the upper crust of these mid-ocean mic reflection data with much improved vertical reso-
ridge structures is clearly quite strong, with seismic Qp lution (shots and receivers placed every 100 m), and
286 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.65 Comparison of Kappus et al., 1995, mean Wide Aperture Profile data with the rock engineering Q c-Vp -depth-porosity-strength
model derived from Barton, 1995, 2002. A UCS value of 100 MPa gives Q  Q c.

Figure 11.66 Waveform inversion results from measurements on the East Pacific Rise near 14°S. Thinner lines show the one standard devi-
ation error band. Collier and Singh, 1998.

applied full wave form inversion to interpret sub-ocean for the initial high value of 0.48. Their results included
crustal structure beneath the East Pacific Rise near 14° S. an estimate of seismic Qp in the low range of 18–30
They provided evidence of extremely high dynamic across layer 2A, with increases at greater depth, as
Poisson’s ratios as high as 0.48, with a sharp drop to shown in Figure 11.66.
0.25 within 200 m of the ocean floor, across the 2A/2B Their calculations suggested a porosity in excess
transition. A very low Vs in the upper 50 to 100 m thick of 30% in layer 2A, which reduced to 6–7% at the
layer-2A (Vp 1.9 km/s, Vs  0.4 km/s) was responsible top of the 2A/2B transition, and further reduced to
Velocity structure of the earth’s crust 287

Table 11.6 Vp-depth data as a basis for rock quality Qc estimation.


From Figure 11.66 data.

Depth below
seafloor Vp (approx.) Qc (approx.) Gradient (s1)

10 m (nominal) 2.0 0.04 –


75 3.0 0.04 15
150 4.0 0.2 13
250 5.0 3 10
800 5.5 4 2

about 5% at a depth of 600 m below the seafloor,


within layer 2B.
Hydrothermal alteration seems to be mostly responsi-
ble for the reducing bulk porosity and for preferentially
(a)
sealing low aspect-ratio cracks. They also cited the higher
porosity of extrusives, i.e. pillow lavas, compared to the
deeper intrusives, which consist of dikes and sills. The
increasing lithostatic (effective) pressure also preferen-
tially closes the cracks or joints with lowest aspect-ratios,
which became predominant across the 2A/2B transition.
As they and others have emphasised, this can have a dra-
matic initial effect on Vp gradients, until ‘seismic closure’
is achieved at sufficiently high effective stress. Table 11.6
gives the approximate indications of Vp-depth gradients.
Referring to Figures 11.44 and 11.65, we see that, in
rock engineering terms we need to ‘curve-jump’ from
initially very low rock quality Q c-values, consistent
with extremely fractured and altered conditions near
the surface, to a typical poor quality jointed medium
(Q c  3 to 4) at greater depth in layer 2B. (b)

Figure 11.67 Upper crustal velocities in the Pacific and Atlantic


oceans for a) 0–150 m.y. b) detail of first 15 m.y.
11.6 Age effects summary for Typical layer 2B velocities (5.2 / 0.4 km/s) shown
Atlantic Ridge and Pacific Rise by dark band, from Houtz and Ewing, 1976 and
Houtz, 1976. (OBS  ocean bottom seismometer,
OBH  ocean bottom hydrophone, log  downhole,
Finally, this chapter will conclude with a broad review
borehole  borehole seismic surveys). Carlson, 1998
of age effects for both mid-Atlantic Ridge and Pacific
and his numerous cited authors.
Rise data. A very wide ranging assembly of seismic vel-
ocity data for uppermost oceanic crust (layer 2A), by
Carlson, 1998, (with one and a half pages of referred
authors) suggested that most of the age-dependent (shown by dark band). Layer 2A appears to persist as a
increase in seismic velocities occurred ‘rapidly’ with vel- low velocity capping of the ocean crust, even when
ocities nearly doubling in 10 million years. 15 m.y. old, as shown in Figure 11.68.
The apparently rather heterogeneous sets of data The trend for increased velocities as age increases is also
(Figure 11.67) when synthesised by Carlson, 1998, using shown clearly by the statistics for 1 m.y., 1–5 m.y.
mean and 9-point interval median values, showed much and 5 m.y. in Figure 11.69, also from Carlson, 1998.
clearer trends of accelerated velocity increase with early The link between hydrothermal alteration and seismic
age, and clear distinction from typical layer 2B velocities velocity increase, due to deposition of minerals first in
288 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.68 Mean velocities (black circles), and 9-point median


velocities (open circles), as a function of median
basement age. Mean velocities are from age intervals
0–1, 1–5 and 5–20 Ma. The data demonstrates that
layer 2A persists as a low velocity (and low rock qual-
ity) layer, capping the crust. Hatched area is typical
layer 2B. Carlson, 1998 and his numerous cited
authors.

the thinnest cracks and joints, gives logical support to the


notion of most rapid alteration when the hydrothermal
activity has been most vigorous.
However, a potential problem of interpretation exists
as pointed out by Carlson, 1998, because hydrothermal
void filling causes a simultaneous increase in velocity and
reduction in hydraulic conductivity, therefore supposedly Figure 11.69 Distribution of seismic velocity in the upper crust
reducing heat flow to the ocean flow. Yet the heat flow (layer 2A) from age intervals 0–1, 1–5, 5 Ma. By
has continued for far longer than the early period of comparison, layer 2B velocities generally range from
velocity increase. When one considers the cubic flow 4.5 to 5.6 km/s. Carlson, 1998 and his numerous
law, suggesting that mineral sealing of small apertures cited authors.
will have less effect on hydraulic conductivity than seal-
ing of large apertures, it seems reasonable to envisage
that the sealing of smallest apertures first will not com- of 2.9 and 4.3 km/s, respectively. Over the last 8.5 m.y.
promise the increase in seismic velocity and the main- there has been local ridge spreading at an average
tenance of a reasonably high heat flow. annual rate of 85 mm/year in this area.
Further interesting details of hydrothermal circula- Their forward modelling of 17 (split) profiles on
tion and mineralization, as an explanation for seismic 0.5 to 8.3 m.y.-old crust for three age ranges is shown
velocity increases with age, were given by Grevmayer in Figure 11.70. When velocity is plotted versus age
et al., 1999. Their work was concentrated in Layer 2A, as in Figure 11.71, a rapid then gradually slowing rise
on the eastern flank of the East Pacific Rise at 14°S, in velocity is seen, which is similar of course, to the
along a 720 km by 25–40 km wide corridor, with only wider-reaching review of Carlson, 1998, shown in sum-
thinly sedimented seafloor of up to 8.5 Ma age. For mary in Figure 11.68. Grevemeyer et al., data indicate
0.5 Ma and 8 Ma crust, they derived P-wave velocities a continuous decrease of the velocity gradient in layer
Velocity structure of the earth’s crust 289

Figure 11.70 Vp-depth models of upper oceanic crust from inver- Figure 11.71 Layer 2A velocity as a function of age (at top of
sion of 17 OBH split profiles, on 0.5 to 8.3 m.y. layer). New results (star-symbols) and various results
crust, from the East Pacific Rise at 14°S. Note the from referred studies, including Carlson, 1998
three age groupings. Grevemeyer et al., 1999. (solid circles). Grevemeyer et al., 1999.

2A, with age up to 10 m.y. Their data only just reaches


strength. These joint-grouting property-improvement
the plateau seen in Figure 11.68, beyond about 8 m.y.
aspects will be described at the end of this chapter.
Grevemeyer et al., 1999, argued convincingly that
hydrothermal mineral filling of open void spaces was
the reason for age dependent velocities, and that veloc- 11.6.1 Decline of hydrothermal
ities in layer 2A remained constant in crust older than circulation with age and
10 m.y. An 8 m.y. crust at the top of layer 2A, showed sediment cover
4.3 km/s which is high for the Pacific. Even 110 m.y.
crust and older, usually showed between 4.0 and It is well established that hydrothermal circulation natu-
4.3 km/s. rally declines with the age of the crust. Anderson et al.,
Interesting parallels to the above hydrothermal mineral 1985, and Evans, 1994, deduced on-axis (zero age)
filling of fractures, can be gleaned from civil engineering, upper crustal permeability of about 6  1012 m2
where the sealing of jointed rock by pre-grouting with (6  105 m/s), this decreasing to about 7  1014 m2
fine-grained micro-cements ahead of tunnels, or the use (7  107 m/s) within 6 m.y. Seismic velocities for
of industrial (coarser grain size) cements in dam foun- crust of the same age were 2.2 km/s and 4.0 km/s accord-
dations, are common ways of both sealing and improving ing to the Grevemeyer et al., 1999, data. The same authors
properties. Quadros and Correa Filho, 1995, measured surmised that permeability may reduce to about 1014 m2
rotations (and magnitude reductions) of the three per- (107 m/s) or less, by the time the crust is old enough
meability tensors, when conducting multiple-borehole to have reached a 4.3 km/s ‘plateau’.
3D hydrotomography before and after grouting. (Holes In thinly sedimented areas, sealing/plugging of crustal
were redrilled after grouting, for the second round of pore spaces appears to extend for 7 m.y., perhaps up to
permeability testing). 15 m.y. However, in regions with significant sediment
These tensor or principal value rotations are inter- cover, the previously open seawater convection cooling
preted as due to sealing of the most permeable and least system is hindered, and temperatures rise, thereby accel-
stressed joint sets. This process can also be interpreted erating the formation of secondary minerals and porosity
by small changes in five or six of the Q-value parameters sealing.
(ratings in Appendix A), which has been shown by Rohr, 1994, used these arguments to explain
Barton, 2002, to cause some dramatic potential improve- 4.3 km/s velocities in only 1.5 to 2.0 m.y. crust at the
ments in the rock mass properties such as modulus of Jan de Fuca Ridge. So Grevemeyer et al., 1999 con-
deformation, seismic velocity, and frictional and cohesive cluded that basement temperature, which is a function
290 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.72 The rock quality Vp-depth-Qc curves with rule-of-thumb estimates for the approximate permeability (m/s) caused by the ‘soft’
porosity (i.e. the jointing). As in the case of pre-grouting with micro-cements ahead of tunnels, one can argue for improve-
ment of various Q-parameter ratings through sealing-mineralization, provided that clay and layer-lattice minerals are not
involved. Harder minerals will have cumulative positive effects on Q-values, and therefore on permeability and other linked
rock engineering parameters. ‘Curve-jumping’ or ‘Q-jumping’ can therefore occur. Based on Barton, 2002.

of heat flow, sediment thickness and sediment perme- along the Q or Qc ‘isoquality’ curves. It is relatively easy
ability, governed the evolution of the seismic properties to invoke this ‘jointed rock’ model as an analogue to
of upper ocean crust. explain a surface or near-surface permeability, as referred
Since massive, non-vesicular basalt has a P-wave veloc- above, of about 105 m/s (1012 m2), with increases to
ity of about 7 km/s (Wilkens et al., 1988), then mature 106 and 107 m/s (1013 and 1014 m2), as age allows
sealed ocean crust, with velocities nearer 4 km/s, must the effective rock mass quality to rise, due to mineraliza-
have a residual porosity consisting of vesicules, fractures, tion sealing in some of the ‘soft’ porosity. At the nomi-
breccia and inter-pillow voids. Most importantly, it is nal 25 m deep ‘surface’ drawn in Figure 11.72, Qc
actually under a low effective stress, just as at the earth’s changing from 0.01 through 0.1 to 1 (i.e. ‘curve jump-
surface, despite (and because of ) 30 to 40 MPa hydro- ing’ due to improving quality with age) implies permea-
static pressure, and because of the shallow sediment and bility reducing from 105 through 106 to 107 m/s.
rock cover. We can express these analogues for rock mass quality
In the rock quality world of Q and Qc (Qc  improvement with age, due to sealing-mineralization
Q  c/100) there is a rough rule-of-thumb (Barton, (as opposed to clay-lubrication) in another way in the
2002) that due to different degrees of ‘soft’ porosity Vp-Qc-M-L interaction nomogram in Figure 11.73.
caused by jointing, the Lugeon value (1L  107 m/s Here we have marked the above analogue for perme-
or 1014 m/s) is approximately inversely proportional to ability reduction with age (roughly 105 to 106 to
Qc. This applies in a central range of rock quality, uncom- 107 m/s) at the nominal near-surface depth of 25 m
plicated by clay sealing of joints, which reduces both Q (central diagonal), and at 50 m. Also shown is the implied
and permeability together, thereby defeating the inverse improvement of modulus of (static) deformation for
proportionality. the rock mass: 2 to 5 to 10 GPa, based on the simple
In Figure 11.72, shown earlier in connection with empirical relation (units: GPa):
velocity-depth gradients, the above rule-of-thumb per-
meabilities have been marked at appropriate depths M(or E mass )  10Q 1/3
c (11.12)
Velocity structure of the earth’s crust 291

Figure 11.73 The rock quality Qc-Vp-M-L interaction nomogram, with appended ‘circles’ to mark the engineering consequences of ‘curve-
jumping’ on permeability, modulus and velocity. After Barton, 2002.

11.6.2 The analogy of pre-grouting need for rock support (bolting and shotcreting quantities
as a form of mineralization are reduced), and of course reduced or negligible inflows.
Permeability K (m/s) may reduce from 105 to 108,
In this final section of Chapter 11, a possible rock engin- 106 to 108 or 107 to 108 m/s, the relative degree of
eering analogy to the hydrothermal mineral sealing of improvement being related to the ‘severity’ of pre-treat-
ocean-floor basalts will be demonstrated, using the ment conditions. Higher pressure injection, from 5 to
analogy of high pressure (5 to 10 MPa) pre-injection 10 MPa excess pressures (above joint water pressure), can
pressures used in an ‘umbrella’ of numerous (20 to 40) cause permeabilities to reduce to between 108 and
boreholes, which are commonly drilled and injected 109 m/s. There is an interpreted 1 to 5 litres of grout per
ahead of leaking tunnels, or where there is environmen- 1 m3 of (locally-injected) rock mass in successful, high-
tal sensitivity at the surface and significant inflows to pressure grouting (Barton, 2004a). In other words there is
the tunnel cannot be allowed. the effect of some joint deformation close to the injection
The argument for improved rock mass quality due holes, made permanent by the subsequent hardening.
to the sealing (by micro-cement and micro-silica) of Such would also be expected during hydrothermal
successive joint sets, is based on the evaluation of individ- episodes of injection, the subsequent ‘fossil’ mineral fill-
ual Q-parameter descriptions and ratings. Each small ings therefore somewhat exaggerating the pre-injection
improvement, like the reduction of the number of effect- apertures and permeabilities, despite some subsequent
ive joint sets due to sealing (i.e. Jn reduces, see Appendix pressure adjustment prior to crystallization. Sealing of
A), may have cumulative effects on rock qualities Q and major channels is ‘a problem’ in both scenarios, because
Qc. There exists in situ proof of improved qualities due to any continued flow will tend to hinder crystallization/
observations of improved stability, less deformation, less hardening in the two processes.
292 Rock quality, seismic velocity, attenuation and anisotropy

Figure 11.74 Three-dimensional permeability monitoring (hydro-tomography) performed before and after grouting, showing rotation of
permeability tensors, and reduction in principal magnitudes. Quadros and Correa Filho, 1995; Barton and Quadros, 2003.
Velocity structure of the earth’s crust 293

Table 11.7 A hypothetical model of potential Q-parameter improvements as a result of both pre-injection with micro-
cement (in tunnelling) and through (hard, resistant) mineral sealing in sub-ocean crust.

Improvement of rock mass properties with pre-grouting, as analogy to mineralization

Effective RQD Increases e.g. 30 to 50%


Effective Jn Reduces e.g. 9 to 6
Jr Increases e.g. 1 to 2 (changed set)*
Ja Reduces e.g. 2 to 1 (changed set)*
Jw Increases e.g. 0.5 to 0.66 (perhaps Jw  1 is achieved)

SRF would reduce only near surface, e.g. 2.5 to 1


(*it may be appropriate to qualify with the word ‘perhaps’ in these cases)

30 1 0.5
Before pre-grouting Q     0.3 (i.e. prior to mineralization)
9 2 2.5

50 2 0.66
After pre-grouting Q     11 (i.e. after mineralization)
6 1 1

Before pre-grouting After pre-grouting Empirical equation

RQD J J
Q   r  w
Q  0.3 Q  11 Jn Ja SRF
c
Qc  Q 
100

Vp  3.0 km/s Vp  4.5 km/s Vp  3.5  log Qc km/s


L  3 (3  107 m/s) L  0.1 (108 m/s) L  1/Qc

M  7 GPa M  22 GPa M  10Q 1/3


C
GPa

As referred to in Chapter 9, it has also been shown by joint sets that may be sealed. More likely, the three sets
Quadros and Correa Filho, 1995 and Barton and represented by Jn  9, could be reduced to two sets
Quadros, 2003 that permeability tensors can rotate and (Jn  4), or even to one set (Jn  2). On the other hand,
reduce in magnitude as a result of grouting. This is shown the grout is an ‘inferior’ fill in relation to hard rock, and
in Figure 11.74. It appears to be evidence of successive may not then provide the expected increases in velocity
sealing of the joint sets, starting presumably with the set and modulus, shown by the empirical equations in Table
that is under least normal stress (probably nearly paral- 11.7. So a greatly increased apparent Q-value, would not
lel to max), or with the most permeable set (or sets) per- then give realistic velocity improvements.
haps caused by (conjugate) shearing and dilation. This The exception would be the case of injection into
presumably could also govern the chronological order of weak rocks with comparable compressive strengths and
deposition in hydrothermally opened fractures and major densities (roughly 20–30 MPa and 1.5 gm/cm3). A beau-
inter-connected pore space in the sub-ocean basalts. tiful example of an igneous intrusion into what was, or
The mechanism by which rock mass quality and rock has since become a weaker rock is shown in Figure 11.75.
mass velocity increases, as a result of both successful pre- This could act as a reminder that the injected hydrother-
grouting and ‘successful’ mineralization of sub-ocean mal fluids, and of course magma, if subsequently very
crust (as it gets hotter due to sediment sealing), could be stiff compared to the surrounding rock, will tend to be
as in Table 11.7 (see Appendix A for descriptions and jointed due both to cooling and deformation, thereby
ratings of the six parameters). maintaing a level of permeability at depth.
The hypothesised improvements in Q-parameters are
very conservative particularly concerning the number of
294 Rock quality, seismic velocity, attenuation and anisotropy

Figure. 11.75 An igneous intrusion (dike) tends to have elevated permeability due to the number of joint sets (typically four: Jn  15). High-
modulus mineralized veins may be fractured by subsequent tectonic deformation, helping to maintain some permeability
despite the ‘sealing’ process. (see Plate 2)
12 Rock stress, pore pressure,
borehole stability and
sonic logging

By their very nature, hydrocarbon-bearing rocks rely on to stress-fracturing, increased permeability, and acceler-
pore-space and permeability for the possibility of hav- ated mud-filtrate invasion. The frequent development
ing recoverable reserves that can be produced at a well. of a near-wellbore, tangentially-distributed discontin-
The necessary entrapment beneath a shale-sealed anti- uum, in cases of insufficiently mud-supported rocks,
cline or by the juxtaposed impermeable layer caused by may perhaps have been overlooked, when modelling
fortuitous faulting, are two basic scenarios for the present- infiltration.
day existence of the reserve. The necessary migration of
the hydrocarbons from source rocks into potential
entrapment structures, without escape to the atmosphere, 12.1 Pore pressure, over-pressure,
adds to the adverse statistics of hydrocarbon explor- and minimum stress
ation. Too close to the surface the sealing properties of
shale, salt or clay-smear in faults, may have been com- Before addressing the details of seismic signatures of
promised by lack of plasticity and too high permeabil- reservoir rocks, caused by numerous environmental
ity. Too deep, the pore space and permeability of the effects such as pressure, temperature and fluid type in
reservoir may be compromised, giving a reduced reserve Chapter 13 (at laboratory rock physics scale) and in
and the need for permeability enhancement and gradi- Chapters 14 and 15 (at reservoir-scale), it is appropriate
ent enhancement, or a decision for non-development. to consider the components and modifiers of the most
Following seismic interpretation of potential fundamental of reservoir parameters, namely the effect-
hydrocarbon-bearing structures, the need for expensive ive stress magnitude. The rock stress and its variations
exploratory drilling and well-testing follows. Besides with direction, depth and location, and the pore pres-
reservoir access for production testing, the hole is used sure and sometimes over-pressure which are influenced
for sonic logging and selected side-wall and regular core by compaction and also by fluid type, are the major
recovery, to better define the properties of the different boundary conditions. Their relative magnitudes affect
lithologies, seals and reservoir rocks. Fortunately, or both the laboratory test simulations, the drilling pro-
unfortunately according to viewpoint and tools avail- gramme, the production planning, and the reservoir
able, rock reacts to the drilling of boreholes with a com- production and depletion, possibly for 50 years or
plex coupling of rock stress and strength magnitudes more in a large reservoir.
(plus the anisotropies of each), not forgetting the neces-
sary subtraction of pore pressure for conversion of the
three principal stresses to effective stresses. The appropri- 12.1.1 Pore pressure and over-
ate selection of wellbore ‘temporary support’ in the form pressure and cross-
of mud pressure, using variable mud weight, deter- discipline terms
mines the state of the borehole wall in the different
lithologies, prior to setting and cementing the casing. Without attempting too much detail, since ‘pore pressure
Due to various opinions about an ‘alteration zone’ analyst’ has become a speciality career choice for numer-
around the wells, there is now widespread acceptance ous petroleum engineers, it is worthwhile following the
of the need for logging while drilling (LWD) with helpful philosophy of Bruce and Bowers, 2002, and men-
monopole and dipole tools, to obtain ‘early’ velocity tioning various cross-discipline differences for describing
responses, which may differ significantly from subse- the effects of pore pressure and over-pressure. Reproduc-
quent wireline logging. The differences are probably due ing selected diagrams from their practical article, we see
296 Rock quality, seismic velocity, attenuation and anisotropy

the following details in Figure 12.1a to d, and Figure 12.2


a to d.
Most fundamentally, over-pressure and the presence
of petroleum products, both increase the pore pressure.
So the effective stress is reduced, which will have the
effect of causing a reduction in P-wave velocity, as illus-
trated in Figure 12.3 from Bowers, 2002.
Over-pressure commonly occurs where low permea-
bility layers (as in shale) prevent fluid from escaping as
rapidly as pore space compacts. Excess pressure in rela-
tion to hydrostatic then builds up as newly deposited
sediments cause squeezing of the trapped pore fluids,
which could be water, oil or gas or even two or three of
these close together.
To conclude this brief section on over-pressure effects,
we may refer to more sophisticated considerations that
over-lap with environmental effects discussed in the next
chapter. Carcione and Gangi, 2000, reported results
from their modelling of seismic attributes of gas gener-
ation and over-pressure. Their model for basin-evolution
showed that pore pressure effects were seismically
visible when the effective pressure was less than about
15 MPa, with oil-to-gas conversion of only 2.5%. Here
they differentiated live oil containing gas, from the
dead oil-free gas which may become seismically visible.
They found that a small conversion of oil to gas was
sufficient to make the pore pressure equal to the con-
fining pressure. The large changes of predicted velocity,
as shown in Figure 12.4, were the result of the fact that
the dry rock moduli were strongly affected by low effect-
ive pressures.

12.1.2 Minimum stress and


mud-weight

A complication for drillers when drilling through inter-


bedded lithologies, or where hole stability is marginal
due to limited rock strength or local ‘structure’, is their
Figure 12.1 a) The basic gradients when there is over-pressure, con-
sidering min and the possibility of failure (hydraulic
fracturing) by too high mud pressure. b) Depth-pres-
sure gradients, showing equivalent mud-weights (in
American units of lbs/gallon), where sea water  8.54
lbs/gallon. [1 lb/gal.  0.0519 psi/ft] c) Change from
pressure-depth axis, to mud-weight-depth format (as
preferred by drillers for obvious reasons). d) To prevent
hydraulic fracturing by high mud-weights, as needed
where there is overpressure, casing will be set to protect
the overlying units from fracturing. Bruce and Bowers,
2002.
Rock stress, pore pressure, borehole stability and sonic logging 297

choice of mud weight. This heavy fluid acts as a tempor-


ary support of the walls of the well during drilling, until
replaced by steel casing that is grouted into intimate,
impermeable contact to the rock. Well testing, possible
minimum rock stress estimation by mini-hydrofracing
and subsequent production occurs through shaped-
charge perforations of the casing in the reservoir inter-
vals, and also in the cap-rock, if minimum stress is to be
measured, to determine the minimum stress difference.
Inter-bedded ‘brittle’ layers (like sandstone as opposed
to ‘sand’), and ‘plastic’ layers (like shale or salt) will likely
exhibit fluctuating minimum principal stress. Since the
shale and salt-rocks may have insufficient shear
strength to tolerate a significant stress difference, the
minimum rock stress (the minimum of three principal
rock stress directions) in the shale (or salt) will often
exceed the minimum stress in the reservoir sandstone
by up to several MPa. An example from measurements
described by Barton, 1986, is shown in Figure 12.5.
The shale (and the salt) may, if encountered at suffi-
cient depth during drilling, require the support of an
active mud-weight to prevent creep or squeezing, as
graphically illustrated by Bradley, 1978, for the case of
deviated wells, in Figure 12.6. So the drillers choice of
mud-weight, or the setting of casing, becomes critical
where support of the well is needed adjacent to a reser-
voir rock like sandstone or fractured limestone or chalk,
which would tend to have a minimum rock stress less
than that of these weaker, sealing ‘plastic’ layers. The
reservoir horizons could potentially fracture, or have a
permeable joint under lower normal stress than the
mud-weight needed to keep the plastic materials from
squeezing and jamming the drill-string. Invasion of
mud (lost circulation), into any reservoir horizon is
obviously very undesirable.

12.2 Stress anisotropy and its


intolerance by weak rock

The local measured variations of minimum principal


Figure 12.2 a) Overpressure in relation to normal hydrostatic pres- stress shown in Figure 12.5 are due to intolerance of
sure and its potential effect on effective stress-depth
stress difference in the weaker shale, as compared to the
trends (in relation to vertical). b) Resistivity, velocity,
sandstone. As discussed earlier, such a mechanism also
density, depth trends with various degrees of deviation
possible where over-pressure changes the effective stress.
occurs in the case of salt. Swolfs, 1977 provided a
c) Pressure at well B is the sum of normal (hydrostatic) graphic comparison of tolerance of stress difference
pressure, plus over-pressure, plus a buoyancy effect ( v  h min) for different rock types. Figure 12.7
caused by the reduced density of the petroleum. d) The shows data for v (overburden, calculated) minus h min
pressure components given in terms of depth-pressure (minimum horizontal principal stress) versus vertical
gradients. Bruce and Bowers, 2002. stress. Rocks such as granite, stronger sandstones and
298 Rock quality, seismic velocity, attenuation and anisotropy

Figure 12.3 Example of high-pressure well from Bowers, 2002, where velocity (and resistivity) undergo reversals, due to ‘under
compaction’.

Figure 12.5 Minimum principal (rock) stress in inter-bedded


reservoir shales and sandstones, as recorded by mini-
hydrofracing tests, to determine optimal stimulation
by MHF (massive hydraulic fracturing), Barton, 1986.

stronger shales tolerate differential stress much better


than weaker shales and salt rocks.
Swolfs, 1977 compilation of vertical (over-burden)
and minimum stress data for these typical North
American reservoir rocks (and of some harder rocks like
granite and gneiss) to depths of 5 km, shows greatest
stress anisotropy near the surface, and in the harder
rock types. Figure 12.8 shows a range of Ko from 0.35
to 2.9. Bedded salt in particular, plus weak rocks like
Tertiary oil shale and Tertiary tuff showed h min/ v
close to 1.0.
Figure 12.4 A model of low frequency (25 Hz) P- and S-wave From the point of view of stress-azimuth dependent
velocities versus excess pore pressure in the case of anisotropic velocity, the added influence of stress-
deep gas resources. The seismic visibility begins due oriented jointing and the character of these joints will
to the sensitivity of Vp and Vs to reduced effective stress, likely be more important than stress anisotropy effects
especially in the dry state (dotted lines). From Carcione on the matrix alone, due to the nature of ‘soft’ porosity
and Gangi, 2000.
(low aspect ratio) jointing. When shear wave splitting
Rock stress, pore pressure, borehole stability and sonic logging 299

Figure 12.6 The driller’s dilemma: avoidance of lost circulation Figure 12.8 Compilation of North American reservoir (and
(hydraulic fracturing or jacking), with the simultan- harder rock) and h min/ v(Ko) ratios as a function
eous need to support the rock walls of the well before of depth. Swolfs, 1977.
casing is in place. Bradley, 1978.

and joint shearing mechanisms are treated in Chapters


15 and 16, it will be found that the influences of
anisotropic stress on the jointing will be of over-riding
importance, since jointing may not always be parallel to
the major stress direction, as traditionally expected.

12.2.1 Reversal of Ko trends nearer


the surface

As one approaches the surface, inter-bedded rock types


resembling reservoir sequences, as illustrated in Figure
12.9, show the reverse of the previously discussed shear
stress intolerance. Hydraulic fracturing tests reported by
Barton, 1981 and Barton, 1986 generally showed low
Ko ratios ( h min/ v) in the weaker materials like shale
and siltstone, and maxima in the sandstones.
This reversal of Ko trends (at a certain, unknown
depth) if a general effect, may have certain ramifications
when comparing stress-induced velocity anisotropy (and
velocity ‘oscillation’) near-surface and at greater depth,
for example when evaluating the applicability of
Figure 12.7 Different degrees of stress difference (or shear stress) ‘research-borehole’ seismic testing. There could be pos-
tolerance of reservoir rocks, compared to granite. sible consequences for the relative magnitudes of attenu-
Swolfs, 1977. ation, as low Ko (in addition to lower stress levels near
300 Rock quality, seismic velocity, attenuation and anisotropy

Figure 12.9 A general tendency for lower Ko ratios ( h min/ v) in weaker rocks like shale and siltstone, according to hydraulic stress meas-
urements at 100 to 300 m depth. Barton, 1981; Barton, 1986.

the surface) will tend to enhance attenuation, as seen in


Chapter 10, as compared to the reduction of attenu-
ation at depth due to higher stress and Ko values closer
to 1.0 i.e. less anisotropy.
Thus Q p and Q s values could be expected to be
markedly lower and exhibit more anisotropy near-surface
than at depth, as compared to rock sequences that did
not show this Ko reversal. Whether such a trend has
been recognised in seismic signatures, for these specific
reasons, is uncertain.
A possible modulus-related explanation for this Ko
reversal was suggested by Barton, 1986. Figure 12.10
shows imaginary stress-strain curves for loading-unloading
(i.e. sedimentation-erosion) curves for a stiff pseudo-
elastic sandstone and for a non-elastic, hysteretic shale.
Imaginary deep-burial (but-on-the-unloading-curve)
moduli of E2 (shale)  E1 (sandstone) and shallow-
burial (but-on-the-unloading-curve) moduli of E3 (sand-
stone)  E4 (shale) would give an explanation for the
above depth-related Ko reversal. Whether the potentially
different (unloading) moduli at these different depths Figure 12.10 Imaginary stress-strain curves for loading-unloading
would cause a part of an inhomogeneous velocity and (i.e. sedimentation-erosion) curves for a stiff pseudo-
attenuation structure, in addition to that caused by elastic sandstone and for a non-elastic, hysteretic
stress anisotropy and joint structure azimuth variations, shale. Imaginary deep-burial (but-on-the-unloading-
remains an open question. curve) moduli of E2 (shale)  E1 (sandstone) and
Another open question, is perhaps whether the ‘oscil- shallow-burial (but-on-unloading-curve) moduli of
lating’ minimum effective stress ( min – pore pressure) E3 (sandstone)  E4 (shale) would give an explan-
in such alternating reservoir rocks is a significant reason ation for depth-related Ko reversal. Barton, 1986.
for variable attenuation. In theory, it should cause ‘oscil-
lations’ of Vp and Vs over-and-above those caused by
the alternating media, with their different intrinsic prop- bedding planes and joints, could each play their role in
erties like density, porosity, modulus and degree of frac- the velocity-depth and attenuation (intrinsic and scat-
turing or jointing. Possibly all the above cyclic properties tering) behaviour, and as we shall see in Chapter 15,
of c, n% and Ko, and the different compliances of the influence shear wave splitting and anisotropy.
Rock stress, pore pressure, borehole stability and sonic logging 301

12.3 Relevance to logging of


borehole disturbed zone

The hydrocarbon reservoir exploration and production


industry has found that the subject of borehole failure
modes is an important ‘complication’ concerning the
interpretation of sonic-logging of wells. There are now
acoustic dipole and monopole logging devices that can
be used in a logging while drilling LWD mode, that acquire
responses from more than one hundred wave forms, in
order to delineate formation fracturing response, and vir-
gin conditions further from the walls of the wells, before Figure 12.11 Shear and compressional sonic logs in ‘compacting
additional ‘alteration’ has occurred from stress and/or shale’ between 1000 m and 2000 m depth in the South
mud-filtrate invasion, as often seen in subsequent wireline China Sea. Miller et al., 1994. ‘Substantial variation’
logging. about the assumed (central) trends is taken by the
There are possibilities for local velocity (and seismic writer as a possible indication of over-stress in numer-
Q) enhancement due to tangential stress increase in the ous layers that, by the nature of the variation, were pre-
case of competent rock like limestones, or low porosity sumably insufficiently mud-supported.
sandstones. In the case of over-stressed, fracturing (‘dog-
earing’) sections of rock, and especially in the case of 1.6 km depth in steeply bedded quartzites: Barton and
incompetent rocks like shales, reduction of velocity (and Bakhtar, 1983). Velocity variations and permeability
seismic Q) will occur, due to the mini-EDZ that form variations as a function of position and radial depth
as a result of drilling. In the radial direction, stress around the tunnel or shaft, can also be determined,
reduction, which may be azimuthally varying, will tend thereby relating these parameters to eventual stress
to locally reduce the velocity, and thereby also the seis- anisotropy. (See Chapters 7 and 8)
mic Q. It is impossible to support each lithology with Plona et al., 1997, reported high frequency P-wave
the ideal mud-weight, so some suffer the consequences, monitoring around a uniaxially loaded block of sand-
just as occurs in a tunnel where the contractor might be stone containing a borehole, and Figures 7.31 and 7.32
trying to reduce costs by under-supporting. (and Section 7.4) gave indications of the significant
In Part 1, Chapter 7, a range of effects that the EDZ velocity anisotropy (10 to 15%) due to unequal tan-
(excavation disturbed zone) could have on seismic gential stress concentration. They also suggested that
velocity inhomogeneity around tunnels was reviewed. the in situ detection of these azimuth-dependent velo-
Due to the smaller size of wells and the use of mud for city differences could be used to predict principal stress
hole support, the recognition of the behavioural data orientations, at stress levels below those needed for
that can be extracted from anisotropic stress effects on break-out. This was possibly signalling the early progress
small-scale EDZ round wells, seems to have come much of important industry developments in dipole, logging-
later to the petroleum industry, than in tunnel engineer- while-drilling, which will be illustrated later.
ing. This comment does not of course apply to borehole Suspicion that wellbore effects must be important for
ellipticity, a much-used historical indicator of the mini- logging results can be high-lighted by reference to
mum horizontal stress axis. Figure 12.11, taken from the sonic logging of a ‘com-
The comment does apply to what lies behind the pacting shale sequence’ in the South China Sea, reported
ellipticity. In tunnels it is all too easy to see, and occa- by Miller et al., 1994. Should we really believe that the
sionally even be killed by, the effects of structure- ‘substantial variation’ around the trends is a true reflec-
induced wedge release, or stress-fractured ‘lenses’ of rock. tion of different ‘virgin’ lithological properties? Are
We are also able to install multiple-position borehole fluctuations of 500 to 1000 m/s about the mean ‘real’,
extensometers (MPBX), both in tunnels and in shafts, or could they be due to accentuation of velocity changes,
to measure the anisotropic radial-distribution of deform- due to positive (m/s) or negative (m/s) mini-EDZ
ation, thereby giving deformation moduli as a function effects, as discussed, and illustrated above? Are the
of direction and radial depth (e.g. 2.5 to 70 GPa varia- logging tools registering formation data, or wellbore
tion of deformation modulus with radial depth, at effects?
302 Rock quality, seismic velocity, attenuation and anisotropy

There is a strong likelihood that mini-EDZ in the 12.4 Borehole in continuum


weaker, less well mud-supported zones, have reduced, becomes borehole in
radial-dependent velocity, due to failure and deform- local discontinuum
ation in the over-stressed zones. Stronger inter-beds
could show an opposite trend due to tangential stress In mining and rock mechanics, there are well known,
enhancement of the velocities. Mini-EDZ penetrating distinctive differences between stress-induced failure of
several diameters can in fact be detected, and circum- hard dilatant brittle rocks like quartzites and hard sand-
vented by deeper sensing, shear-wave based, dipole log- stones (extensional splitting and crushing), and inter-
ging tools. However the probable discontinuum reality mediate strength and less dilatant rocks, in which the
(caused by log-spiral shearing, to be illustrated shortly), traditional ‘dog-earing’ takes on a different shape due to
is only referred to as e.g. ‘shale alteration’. There is also log-spiral shear failure. These two basic modes were
the question of fabric and jointing and bedding plane contrasted by a person with a broad experience of civil
effects, not least on the progress rates for mud-filtrate and petroleum related fields; Maury, 1987 (ISRM com-
invasion. This seems not to be addressed, as yet. mission on failure modes). The two basic modes were
Fjæer and Holt, 1999 addressed the possible effects linked to possible dilatant/non-dilatant rock failure
of the theoretical, isotropic, elastic stress distribution behaviours by Barton, 1987.
around a wellbore on sonic logging results. They sug- In Figure 12.12, various modes of continuum failure
gested that with conventional long-spaced sonic tools, (of the previously intact matrix) have been assembled
the zone of changed velocity due to the reduction in for comparison. They include a recently reported ‘tabu-
radial stress in relation to back-ground stress, would be lar slot’ type of breakout, probably related with com-
mostly within a 1⁄2 well radius, and would not be detected, paction band development during laboratory-scale
unless specially designed tools were used. They thought drilling in porous blocks of sandstone.
that the larger (and possibly also smaller) P-wave in the The more plastic model material, driven to the extreme
tangential directions would not be measurable by a con- of ‘well closure’ conditions, showed undrained strength
ventional sonic logging tool. mobilization and residual strength development reach-
However, the authors referred to Plona et al., 1998 ing to many diameters. There was also a compacted zone.
who had suggested use of a multi-pole logging tool One may speculate that subsequent recovery (re-drilling)
capable of distinguishing between velocities in different of such a zone would still leave permanent effects in rela-
azimuthal directions, close to the well. Plona and col- tion to velocity reduction and seismic Q reduction.
leagues work, already referred to in Part I, is a central Local, dramatic changes of seismic properties should not
rock mechanics-based part of the newer well-logging deceive those interpreting well logging.
developments that will be described shortly. The variable azimuth drilling described by Addis
Concerning processes important for well logging et al., 1990 was performed with the important three-
interpretation, we must consider modes of well behav- dimensional ‘detail’ of drilling while under 3D stress
iour, going strictly beyond elastic isotropic stress- states, as happens in reality. The same result is not
distribution concepts. As we shall see shortly, the evidence obtained when loading a block with a pre-drilled hole.
to be presented for deeper penetration of mini-EDZ, The deep log-spiral development in the weak (0.5 MPa)
presumably related with the ‘shale alteration’ referred to cemented-sand tested by Addis et al., 1990, typically
by Brie et al., 1998, is much in line with current indus- occurred when the major principal stress was about ten-
try interest. There turns out to be a serious potential times higher than the uniaxial strength, with the minor
contrast in logging results, when comparing the later and intermediate principal stresses of 60% or 80% of the
(e.g. 1–2 weeks later) result of wireline logging, with maximum.
the few hours delay represented by LWD, or logging The borehole failure at the bottom of Figure 12.12 is
while drilling. probably a typical mode for not just non-dilatant,
The more recent shear-wave anisotropy based logging, but actually contracting-with-shear clay-rich materials.
images a volume of up to several borehole diameters Encountered in tunnelling situations, such fault-related,
away from the wall, also beyond the stress-related frac- clay-rich materials may flow to form a weakly-inclined
turing and mud-filtrate invasion or ‘shale alteration’, ‘delta’, flowing many tens of meters from the previous
thereby giving presumed ‘virgin’ formation attributes tunnel face (e.g. a Nathpa Jakri HEP headrace tunnel
as well. access drift, NE India). A case of 7,000 m3 of fault-zone
Rock stress, pore pressure, borehole stability and sonic logging 303

Figure 12.12 A wide collection of failure modes for circular openings. Upper pair: from Maury, 1987 (ISRM commission on failure modes).
Second pair: from NGI joint industry project and Addis et al., 1990. Third pair: left - deep log-spirals from NGI study by
Rawlings, 1991, right – slotted breakouts sketched from Haimson, 2003. See Lee and Haimson, 2006 for recent results.
Bottom pair: from NGI joint industry project: Bandis, 1988. Undrained shear strengths (in kPa) from centre: 0–5 (f ), 5–10
(p), 10–30 (p), 70–75 (c), 40–50 (s), and 50–60 (v). (f: flow, p: plastic, c: compacted, s: softening, v: virgin). See Zoback
et al. 2003 for detailed theoretical analysis of break-out phenomena.
304 Rock quality, seismic velocity, attenuation and anisotropy

Table 12.1 Failure and deformation modes typical in tunnels (after Barton, 2004b) may also apply at greatly reduced scale to the mini-EDZ
that are likely to occur locally, where insufficient mud-weight has been applied.

Description Mode of behaviour

1. Hard, massive, brittle rocks that dilate during failure even Extension failure, thin-walled stress-slabbing, dynamic ejection,
when stresses are high. Stress-induced failure may be delayed bursting. The symmetric ‘dog-eared’ fall-out due to the anisotropic
as ‘strength corrosion’ occurs. stresses may have a ‘nose’.
2. Hard or medium hard, bedded and jointed rock that can Anisotropic response. Shear stress dissipates by slight shear on
shear and dilate along structural planes, while under moderate bedding planes and joints. Deformations are moderate. Block falls
to high stress levels. can occur.
3. Soft, massive, non-brittle rocks that may, or may not dilate Failure may occur by log-spiral shear development and tangential
during shear failure. Typical for young e.g. Tertiary rocks such strain. Radial deformations are large, and pressure on support is
as the mudstones and siltstones in Japan. high. Multi-diameter influence.*
4. Very soft, plastic rocks (and clays) that contract when sheared Post peak strength loss reaches an extreme of virtual ‘flow’, with
under significant stress levels. in-rush of failed material.

(*Twin tunnels need pillars 4 to 5 times their span c.f. Japan, Taiwan, in order to minimise excavation interaction)

debris, filling a 70 m2 tunnel for 100 m has also been terms it is a case of ‘c then tan ’, not ‘c  tan ’.
seen by the writer, in the Pinglin tunnels in Taiwan. Numerical models that are programmed, or manually-
Water pressures may of course be fundamental catalysts steered, to dissipate cohesion while mobilizing friction
for such dramatic and sometimes tragic events. are capable of matching observed behaviour. The BEM
The ‘slotted’ failures described by Haimson, 2003, (Boundary Element) fracture mechanics code FRACOD
were obtained in polyaxially-loaded blocks of Berea (Shen et al., 2002) seems to model log-spiral fracture
sandstone, with 23 mm diameter hole-drilling while development, and over-stress dissipation, in a particu-
under stress, parallel to the intermediate vertical stress. larly realistic way.
Haimson’s samples with 17% porosity developed ‘con- If stress anisotropy (and stress magnitudes in relation
ventional’ dog-earing, while the weaker 22% and 25% to rock strengths) are sufficient to cause break-out, then
porosity samples developed increasing lengths of sym- clearly the principal horizontal stress direction is easy to
metric ‘slotting’ (see simple holes-with-slots traced from predict from calliper logs, it being at right-angles to the
the experimental results, reproduced in Figure 12.12). The largest well diameter in the case of a vertical well. This
uniaxial strengths of the three sandstones were 53 MPa, ‘simple’ situation may be disturbed by break-out caused
42 MPa and 22 MPa respectively. The depicted ‘slotting’ by structure, i.e. steeply dipping jointing causing wedge-
occurred during drilling, with principal applied bound- shaped fall-outs assisted by local bedding.
ary stresses of 50, 60 and 90 MPa, i.e. up to four-times Such mechanisms are a sign of a mini-EDZ that is in
the uniaxial strength of the weakest sandstone. progress around the well, and one that could be used in
Table 12.1, from Barton, 2004b, summarizes a quite dedicated, azimuthal, short-base, well-logging. The
complete range of failure and deformation behaviours influence of stress fracturing on velocity anisotropy,
for tunnels, with potential relevance to mini-EDZ devel- including but beyond the pre-failure states investigated
opment around boreholes, and possible consequences for by Plona et al., 1997, means that there will be anisotrop-
sonic logging, and for bed-property enhancement or ically distributed fracturing around eventual over-stressed
degradation, as discussed above. part of a well, that had not been adequately supported
As emphasised by Barton, 2004b, the actual modes by mud pressure.
of physical behaviour experienced by boreholes and To emphasise (and probably exaggerate) the likeli-
tunnels, are unlikely to be predictable when modelling hood of possible serious differences in EDZ-potential
with conventional Mohr-Coulomb type (c  tan ) down a well in alternating hard/soft rocks, one may refer
shear strength criteria, because rock tends to fail first again to some of the borehole stability studies reported
by loss of continuity at small strain, caused by loss of by Addis et al., 1990. These were performed in the late
local tensile or cohesive strength, followed by frictional 1980s at the Norwegian Geotechnical Institute, in a
mobilization at larger strain. In modified Mohr-Coulomb joint oil-industry study. Figure 12.13 shows one of the
Rock stress, pore pressure, borehole stability and sonic logging 305

Figure 12.14 Closed-form elasto-plastic analysis of an internally-


supported well, compared to the physical-model
reality of log-spiral shear-failure surfaces, that help
to dissipate the highest stresses further into the rock
mass. NGI contract report, 1990.

application (via flatjacks) of three equal or unequal


Figure 12.13 One of a series of log-spiral-type failures around principal stresses, with drilling performed into the mod-
heavily stressed model boreholes that were drilled at elled block of rock while under stress, either as a ‘verti-
different deviation angles in relation to a moderate cal’ well, or at different deviation angles with respect to
3D stress anisotropy. Addis et al., 1990, and NGI any of the principal stresses, as shown in this case.
contract report, 1990. Application of an analytical closed-form elasto-
plastic isotropic continuum model to a heavily stressed,
‘symmetric’, double-cusp, intersecting log-spiral modes drilling-mud-supported well, shown in Figure 12.14
of shear failure that was exposed around the borehole, suggested that a narrow, deeply-penetrating, elliptical
following polyurethane-foam hole-stabilization, while ‘plastic’ zone would develop, where the Mohr-Coulomb
still under stress. Cubic 50  50  50 cm blocks of shear strength criterion was exceeded. The reality – as
model sandstone ( c  0.5 MPa) were used for 3D far as a physical model represents reality – was for one
306 Rock quality, seismic velocity, attenuation and anisotropy

of earth science’s favourite continuum theories to be demonstrates a distinct rotation of the dark grey-scale,
ignored, and instead we witnessed the development (in from a ‘NNE-SSW’ to ‘E-W’ azimuth. The grey-scale
every case), of a 1 to 2 diameter wide zone of log-spiral represents F. of S.  1.0.
failure surfaces, making in effect a discontinuum, which
actually allowed the wall-rock to dissipate the highest
shear stresses further into the surrounding rock, just as 12.5 The EDZ caused by joints,
occurs in tunnelling through over-stressed ‘squeezing’ fractures and bedding-planes
rock. A ‘radial’ strain is observed in the opening – there
is measured radial closure – but the reality is tangential Due to the influence of deformation of ‘soft’ as opposed
strain due to the log-spiral shearing. This mechanism to ‘hard’ porosity, a borehole for hydrocarbon explor-
has been well documented by Aydan et al., 1992. ation that penetrates variably jointed and faulted
These three-dimensional failure surfaces ‘followed ground, will actually experience variable small amounts
the borehole’, and also curved in front of the face of the of deformation, due to different degrees of joint closure,
hole, whenever the holes were drilled into the highly joint opening, and joint shearing. There will also be the
stressed blocks of rock simulant. This was seen follow- pseudo-elastic response, due to both loading (at the dia-
ing sectioning. Such an EDZ, developed in an insuffi- metrically-opposite max locations) and potential
ciently mud-supported weakness zone, or in a bed of unloading (at the diametrically-opposite min loca-
shale adjacent to a reservoir sandstone (with insuffi- tions) of the matrix as well as the joints, the latter usually
cient hmin to tolerate higher mud weight), would tend dissipating some of the theoretical (isotropic, elastic)
to locally accentuate both the expected low seismic peaks of maximum and minimum tangential stress.
velocity, and an expected low seismic Q value, that This process will occur even with a constant mud-
could be obtained from seismic logging. weight, since the mud – unlike rock bolts in a tunnel –
Before leaving rock failure mechanisms in the intact cannot prevent joint movements of unequal magnitude
matrix around boreholes, a hybrid (intact plus jointed) at different points around the opening, although the
result will be shown from modelling with the FRACOD mud may help to make them very small.
fracture mechanics BEM code of Shen, described in Shen Figure 12.17 shows four idealized, two-dimensional
et al. 2002. The two models shown in Figure 12.15, per- models of circular hole excavations in anisotropically
formed by Shen for the case of a deep excavation under loaded ‘blocky-rock’, with four successively reduced block
high stress, had deliberately sparse jointing. Different lev- sizes. There are 250, 1000, 4000 and 10,000 blocks in the
els of ‘log-spiral-type failure are demonstrated, depending respective models. Two of the coarsest block geometries
on the ‘disturbance’ to stress distributions, caused by are shown in the top of the figure, while all four of the
jointing, which dissipates some of the highest, near-wall joint-shearing fields are given. The reducing block sizes
tangential stresses. can be glimpsed within the two lower models. These
The resultant ‘dispersed’ fracturing (as opposed to models represent ‘crushed zones’ at borehole scale, i.e.
jointing) would presumably be a strong source of scat- gross approximations to fault-zone-crossing stability
tering attenuation – and velocity reduction. Note that problems.
the principal applied stress levels in the case of this Figure 12.18 demonstrates analytical Mohr-Coulomb
strong rock were only about 35–38% of the uniaxial formulations used by Shen to produce the general result
strengths. A brittle sandstone of 50 MPa UCS would be for zones of shearing on differently inclined, conjugate or
acted on by an equivalent ratio of effective stress beyond perpendicular joint sets. The four-sector EDZ, at roughly
about 1200–1300 m depth, with standard density and 45°, 135°, 225° and 315° is bisected by the two applied
pore pressure assumptions, considering a H max value boundary principal stresses, e.g. a ‘N-S’ H and an ‘E-W’
no larger than the vertical effective stress. h. Shear displacements occur on the joints in these four
Of particular interest to the frequent velocity discrep- sectors, even when stresses are isotropic, as shown by
ancy between LWD and later wireline logging of the UDEC-MC or UDEC-BB modelling of tunnels through
same formation (see later examples), is the possible horizontally and vertically jointed rock.
development of fracturing over time, and the result this A more realistic model of ‘fractured rock’ (such as
has on the ‘factor of safety’ against further shear failure. might be found in the neighbourhood of fault zones),
Figure 12.16 shows two episodes of fracture develop- actually representing slightly random jointing in tuff, is
ment over time from the same FRACOD model, and shown in Figure 12.19. The model represents a bored
Rock stress, pore pressure, borehole stability and sonic logging 307

(a) (a)

(b) (b)

Figure 12.15 Fracture mechanics BE modelling of circular openings Figure 12.16 Fracture development over time in a FRACOD
in high over-burden environments, using the FRA- model of a circular opening in a jointed zone. Note
COD code developed and demonstrated by Shen. The the ‘rotation’ of the diametral-pair of red regions,
modelled rock mass was hard and of high modulus which represent low factors of safety against shear fail-
(UCS  160 MPa, M  50 GPa), with assumed shear ure. Further fracturing dissipates and displaces the
strength of c  40 MPa,   35° and appropriate low F. of S. zones, suggesting that more fracturing
fracture toughnesses at the modelled effective stress could occur across the ‘E-W’ diameter. Changed seis-
levels of 30 to 60 MPa. a) EDZ with sparse jointing mic response over time is easy to imagine, also a mud-
and 1 (‘NS’) of 55 MPa and with 2 (‘EW’) of 40 MPa. filtrate invasion speed that could be highly
b) EDZ with significant jointing and 1 (‘EW’) of non-uniform, due to developing permeability in the
60 MPa and with 2 (‘NS’) of 30 MPa. Note: deform- partly connected discontinuum. Nick Barton &
ations do not track subsequent fall-out/loosening. Associates 2005 contract report. See Figure 12.15
Nick Barton & Associates contract report, 2005. for input data. (See plate 3).
308 Rock quality, seismic velocity, attenuation and anisotropy

Figure 12.17 Shear deformation zones developed around circular openings, as a result of slip along both joint sets, when excavated under
anisotropic stress, and when with limited joint strength, e.g. due to clay-smear or general smoothness. Cundall’s distinct-
element UDEC-MC code (with Mohr-Coulomb sub-routine), was used for the analyses, with four successively reduced block
sizes. Note the increasing multi-diameter EDZ, as block size and therefore rock quality reduces. Line thickness depicts joint
shear magnitude. Shen and Barton, 1997.

Figure 12.18 Theoretical, Mohr-Coulomb based solutions for shear displacement zones
involving slip on conjugate joint sets. Note that the largest joint-shearing
EDZ is for the 60° case, in which shear stress is maximized relative to the
joint directions. The smallest joint-shearing EDZ is for the 90° case, in
which the shear stress is minimized relative to the joint directions. Shen
and Barton, 1997.
Rock stress, pore pressure, borehole stability and sonic logging 309

Figure 12.19 UDEC-BB modelling of a TBM tunnel EDZ in jointed tuff, at a simulated 650 m depth. Note far-field h  fi v. This model
is assumed to represent a generic case of heavily fractured (faulted) and sheared rock close to a vertical borehole. Clockwise
(with 8 m full-scale excavation diameter): principal stresses (max. 30 MPa), deformations (max. 11 mm), joint shearing (max.
8 mm), and joint (and block-corner channel) conducting apertures (max. 2 mm). Hansteen, 1991 (NGI contract report), and
Barton, 2000.

excavation at 650 m depth, and had the predicted, close- each set, as recovered from deep boreholes at the UK
to-the hole EDZ behaviour. (Note that the model’s Nirex Sellafield’s site in NW England. A second model,
loaded boundaries extended some 10-diameters from this time representing a horizontal bored tunnel at
the hole, and the hole was also close to a simulated clay- 400 m depth through inter-bedded sandstone and shale,
bearing fault.) Four basic EDZ components are shown: is shown in Figure 12.20.
If we scale the roughly 5–15 mm range of deform-
● Principal stresses (‘N-S’), and redistributed tangen-
ations from 8 m tunnel size in hard rock, to a nominal,
tial stress
and convenient, 16 cm well size, with correspondingly
● Displacement vectors: maximum parallel to ‘N-S’
reduced rock strengths, the magnitudes become 0.1 to
stress maximum
0.3 mm. The deformations, although small (i.e. sub-
● Joint shearing (clockwise or anticlockwise) and
millimetre size when the hole is stable), are probably up
proportional to line thickness
to a few orders of magnitude larger than the displace-
● Joint (and block corner channel) conducting
ment discontinuities that are sufficient to generate joint
apertures
compliance changes (normal and shear), that are of
The modelled rock was actually jointed tuff, with recognisable (inverted) magnitude, in relation to the
input data from index testing of numerous joints from joint stiffnesses familiar in rock mechanics. The lower
310 Rock quality, seismic velocity, attenuation and anisotropy

dynamic compliances will signal the reduced velocity


and the reduced seismic Q, via the reduced dynamic
(and static?) moduli.
Since both isotropic and anisotropic stresses, but
especially the latter, can (indeed must) cause a mini-
disturbed zone around a well, it is likely that this could
influence dipole sonic logging down the well. In the
case of the sedimentary rock model, with shale beds, at
a moderate 300 m depth, the joint shearing and deform-
ation EDZ are each of several hole-diameters in extent.
Is it not possible that the mini-EDZ illustrated here,
can be the root cause of much of the intense fluctuation
seen in the sonic logging results presented in Figure
12.11? Low Vp and low seismic Q zones could be
depressed further by an accentuated EDZ. Conversely,
high Vp and high seismic Q could be further elevated
by the effect of high tangential stresses close to the
borehole wall, where there was no failure or joint defor-
mation (or joint presence) to speak of.
Consider the following cross-discipline parallels.
When assessing the rock quality of the walls and arches
of tunnels, the observed rock, which is the visible part of
the tunnel-scale EDZ, is classified (using the Q-system:
see Appendix A), in order to select appropriate rock
reinforcement (grouted rock bolts) and tunnel support
(sprayed, steel-fibre-reinforced concrete). The latter is
the equivalent of the borehole mud pressure, and is
badly needed in a complete load-bearing ring, in the
rapidly deteriorating and deforming clay-bearing zones,
in order to control deformation and prevent local tun-
nel collapse.
Outside the tunnel EDZ, the rock mass would be
characterized as a better quality rock mass. If one per-
formed both sonic logging and azimuthal dipole sonic
logging – a development described by Brie et al., 1998
(see later), the borehole mini-EDZ can be classified by
the one tool, while the hydrocarbon-bearing or reser-
voir sealing formation away from the immediate influ-
ence of the hole, can be characterized with higher
velocity by the other tool. The same arguments as above
apply to seismic Q. As we shall see, at least part of this
Figure 12.20 UDEC-BB modelling of a TBM tunnel EDZ at rock mechanics logic is being applied in modern well
300 m simulated depth, in inter-bedded sandstone
logging, with multi-wave-form acquisition.
and shale. Note far-field h  v. This model is
A layered model, consisting of a ‘thinly-bedded’
assumed to represent a generic, shallow horizontal
well. From top: principal stresses (max. 50 MPa),
model sandstone and micaceous inter-layers that was
deformations (max. 15 mm), joint shearing (max. expertly fabricated by Bandis while working at NGI, is
8 mm). Note that ‘buckling’, or cracking of thin beds, shown in Figure 12.16, to emphasise again that if one
does not occur in UDEC, unless the necessary failure starts with a discontinuum, the possibility of deform-
surfaces are discretized before-hand. Chryssanthakis, ation is increased. In contrast to the jointed UDEC
1991 (NGI contract report), and Barton, 2000. models shown earlier, this steeply inclined bedding
Rock stress, pore pressure, borehole stability and sonic logging 311

allowed a buckling mode of deformation to develop, 12.6 Loss of porosity due to


causing an elliptical yielding zone, with a long-axis of extreme depth
several diameters. It is easier to imagine this case when
rotated through 45°. A mud-pressure capable of pre- Loss of porosity with pressure or depth is well docu-
venting the buckling action can readily be imagined mented in sandstones recovered from sedimentary basins
causing unwanted hydraulic fracturing, as depicted in such as the North Sea and the US Gulf Coast. These
Figure 12.6. Such buckling mechanisms are common basins typically have porosity versus depth curves with
in thinly-bedded coal-measure rocks, and presumably mean porosity of about 23 to 27% at 2 km depth, decreas-
can compromise the integrity of horizontal wells, if ing to 3 to 7% at 7 km depth. This change in porosity sig-
drilled where beds are very thin. nals changes (increases) in mechanical strengths and of
In the foregoing, the possible effects of the weakness course increases in seismic velocities, and, all other things
of formations, in the face of drilling with insufficient being equal, would suggest reduced attenuation or higher
mud-pressure, have mainly been addressed. Accompany- values of seismic Q. In a downhole or reservoir environ-
ing potential shear-failure, there is a tendency for dila- ment, inhomogeneity caused by structure, for example
tion, unless stress levels are high enough to suppress this more intense fracturing or jointing in the resulting stiffer
expansion. The local and bulk dilation is of course part sandstone might have a reversed effect on the above
of the ‘alteration’ zone phenomenon, and is a basic rea- assumption of reduced attenuation with depth.
son for an initial velocity reduction (e.g. under LWD), An assessment of the onset of the brittle ductile tran-
followed intuitively by increasing permeability, there- sition of sandstones reported by Scott and Nielsen,
fore accelerated mud-filtration weakening, with a possi- 1991, has relevance here in the case of increased depths
ble spiral into lower velocity territory, by the time of of burial, and possible components of behaviour around
wireline logging. The permeability of a damaged bore- deep wells, in view of the approximate doubling of the
hole wall is unlikely to remain constant in the interval field effective stresses in the tangential stress arch
between LWD and wireline logging. around an initially non-yielding borehole wall, i.e. when
still responding ‘elastically’ with horizontal stresses of sim-
ilar magnitude, giving (max and min)  3 1  3.
The authors Scott and Nielsen, subjected sandstones
with a wide range of porosities (6.0%, 13.9%, 19.9%,
20.9%, 22.0% and 27.6%) to triaxial confinement
(0.1, 5, 10, 30, 50, 70, 90, 110 and 130 MPa), and also
reviewed tests at much higher confining stresses. All
their tested sandstones had similar grain size, shape,
orientation and composition, and all were cored from
fresh unweathered blocks, without strong bedding plane
anisotropy.
Figure 12.22 shows how the higher porosities were
readily driven to the brittle-ductile boundary; solid cir-
cles representing ductile (non-dilatant) cataclastic flow,
while open circles represented brittle, dilatant, shear
fracture.
On the shear strength development at reservoir stress
levels are shown in Figure 12.23 subject of extreme bur-
ial effects, Chuhan and Bjørlykke, 2002 contrasted the
normal mechanical compaction of sandstones down to
Figure 12.21 A buckling mode of deformation when penetrating
2–2.5 km depth, with grain crushing phenomena at
horizontal or steeply dipping thinly-cycled beds, greater depths. The ‘limits’ of normal mechanical com-
would also tend to compromise the ‘correct’ value of paction appears to correspond to the ‘knee’ seen in
logging-based Vp and seismic Q, in the case of dipole velocity-effective stress tests in the laboratory up to about
azimuthal-logging. Model fabricated by Bandis, 25 MPa. They mentioned the precipitation of quartz
1987 (pers. com.). cement at greater depth, where temperature was higher,
312 Rock quality, seismic velocity, attenuation and anisotropy

causing enough strengthening of the grain framework


to prevent further compaction.
According to Chuhan and Bjørlykke, grain fracturing
is commonly seen in North Sea reservoirs, in coarse-
grained sandstones, but rarely in fine-grained sand-
stones. Some deeply buried sandstones (5 km) at
Haltenbanken showed evidence of more extensive grain
crushing, because grain coatings had apparently delayed
quartz cementation. Deeply buried reservoir sand-
stones (6 km) from the Azerbaijan area, that were
subject to rapid subsidence and a low geothermal gradi-
ent had very little quartz cement and showed evidence
of intense grain fracturing. This was assumed to have
occurred when the effective stress was about 40–50 MPa,
at 4–5 km depth.

12.7 Dipole shear-wave logging of


Figure 12.22 High pressure triaxial tests on sandstones of various boreholes
porosities, showing the approximate limit of brittle
behaviour. Scott and Nielsen, 1991 and cited Brie et al., 1998, in a multi-authored, major contribu-
authors. Note the extreme confining pressures. Solid tion to ‘new directions in sonic logging’, provided a fas-
circles represent cataclastic flow, open circles cinating glimpse of the newer techniques of well logging,
represent brittle shear fracture.
and of the challenging conditions that were met in dif-
ferent types of formations. Their article confirms many
of the foregoing suspicions that what we have termed
mini-EDZ, are indeed a source of concern in certain for-
mations, and more importantly, that these deeper ‘alter-
ation zones’ can be detected and seismically characterized,
and avoided, with the help of novel dipole transmitter
tools, that generate flexural waves. Flexural waves are
shear waves that are polarized into fast and slow direc-
tions, and penetrate several hole-diameters into the for-
mation, thereby revealing potential stress-induced
‘alteration’, and/or drilling mud-induced alteration. As
we shall see in much greater detail in Chapter 15, shear
wave measurements seem to be the most useful mode
for fractured reservoir interpretation at many scales.
The amount of sophisticated equipment that well
drillers may now connected to ‘follow’ the drill-bit, using
a down-hole mud motor in the bottom-hole assembly (or
BHA), is impressive for its sophistication and necessarily
compact nature. It resembles in some ways the several
hundreds of meters of generally much less sophisticated
Figure 12.23 Shear strength envelope separating brittle and duc-
back-up that are set to be pulled (on rails), behind a large
tile behaviour for sandstones of various porosities. tunnel boring machine. Consider this paragraph from
Scott and Nielsen, 1991 and cited authors. Solid Brie et al., 1998, concerning an interval of drilling down
symbols represent cataclastic flow, open symbols an Angolan (West African) offshore exploratory well:
represent brittle shear fracture, and half-black ‘A long single-bit run was conducted over a 7-day period
represents transitional behaviour. that covered a depth from 1000 to 8000 ft. The wellbore
Rock stress, pore pressure, borehole stability and sonic logging 313

was deviated 20° in this interval. The BHA consisted of a


PowerPak mud motor, a CDR Compensated Dual
Resistivity tool, a PowerPulse MWD (measurement while
drilling) telemetry system – for real-time transmission –
and an ISONIC sonic-while-drilling tool. The ISONIC
tool, placed above the PowerPulse system, was approxi-
mately 104 ft (32 m) away from the bit. At the average
rate of drilling, the LWD (logging while drilling) meas-
urements were made fewer than four hours after the for-
mation was first cut. Wireline sonic logging was run after
the 7-day drilling run was completed, and then only after
circulating the well for several hours.’
Brie et al., 1998 emphasised that LWD logging while
drilling sonic logs, acquired in freshly drilled rock, show
‘remarkable differences’, compared with wireline meas-
urements that are usually taken many days, or a week or
two, after the drilling has exposed the formation. They
pointed out that both well surveys bring important but
different information about wellbore rock properties.
The cause of these fundamental differences is summar-
ized in a helpful diagram from their publication, shown
in Figure 12.24.
Since the ‘altered zone’ around the borehole may con-
tinue to develop during the week or so that may sepa-
rate the two types of logging, the later wireline log may
be influenced by a reduced modulus in an annular zone
Figure 12.24 Stress-related damage, possibly compounded by
around the borehole, particularly in soft formations.
subsequent water or mud-filtrate weakening, have
Water uptake in this zone (from mud-filtrate), as well as
the potential effect of reducing the modulus in the
the initial over-stress, will also reduce the modulus in several days or one week delay between LWD (logging
shales and shaly sands, according to Brie and co-authors, while drilling a few hours behind the drill-bit), and
lending support to our initial question concerning the the subsequent wireline sonic logging. Entrapment
real meaning of the South China Sea log shown earlier in of wave-fronts in the lower modulus damage zones
Figure 12.11. A so-called ‘bi-compressional arrival’ may results in ‘bicompressional arrivals’: i.e. second
be registered – a phantom arrival too fast to be a shear arrival compressional waves. Brie et al., 1998.
wave – actually caused by trapping of the wave-front by
the low-modulus damage zone.
Figure 12.25 illustrates the novel dipole transmitter
principle. During logging, flexural waves are generated fracturing, with fastest velocity (as also with P-waves),
by sequential firing in two perpendicular directions (see in the direction parallel to structure, possibly also align-
bottom of tool), first along the tool x-axis, then along ing with the major principal stress direction. (In Chapter
the tool y-axis. These induced shear waves are split and 15, numerous examples of this exceptionally fortuitous
polarized into fast (qS1) and slow (qS2) directions, property of S-waves will be given, from fractured reservoir
respectively parallel and perpendicular to dominant analysis). The slow direction (as also with P-waves) is per-
fracturing or to the principal (horizontal) stress direc- pendicular to the fracturing – which could be microc-
tion (if the tool is in a vertical hole, and more distant racks, cracks, joints or faults – according to the scale of
response is analysed). The shear waves are registered by example considered. Shear-wave splitting, and polariza-
the dipole receiver pairs shown at the top of the tool. tion is one of the most valuable of all seismic anisotropy
The important property of shear-wave splitting is properties for fracture and fracture-fluid investigations,
illustrated in Figure 12.26. In the example shown, the perhaps matched by the anisotropic and dispersive atten-
shear wave splitting is caused by the y-axis aligned uation of P- and S-waves (see Chapters 14 and 15).
314 Rock quality, seismic velocity, attenuation and anisotropy

Figure 12.26 An illustration of the shear-wave splitting and polar-


ization phenomenon, from Brie et al., 1998. Note
the longer wave length of the fast shear-wave.

Figure 12.25 Dipole transmitter tool for sequential firing in


x- and y-directions, to generate flexural (shear) waves
that become split and polarised due to dominant are sensitive to properties deeper within the formation
fracturing or stress, thereby becoming fast (with longer than the surface effects of drilling.
wavelength) parallel to structure, and slow (with A significant finding regarding the distinction
shorter wave length) perpendicular to structure. Brie between stress-related anisotropy around boreholes,
et al., 1998. and fracture related anisotropy, is shown in Figure
12.27 from Winkler et al., 1998 (his co-authors Sinha
The two orthogonal dipole transmitters, and the and Plona were also contributors to Brie et al., 1998).
multiple receiver pairs, which are aligned in orthogonal Borehole stress concentrations (in competent rock like
directions, measure the components of slowness in any the tested Berea sandstone) caused radial stress gradients
direction within planes perpendicular to the borehole. that were different in the two principal directions,
In fact four sets of waveforms are recorded. By min- thereby causing frequency dependent (dispersive) effects
imizing (with Alford rotation), the cross-receiver such that the shear-wave that is polarized parallel to the
amplitude-based energies, the rotated direction of the major stress, is fastest only at low frequency (25 Hz).
fastest shear waves becomes the fast-shear tool azimuth. The fact that shear wave anisotropy allows the inves-
A magnetometer relates this direction to true north. tigation of a volume of the formation up to several
The logging tool results are converted to graphic diameters from the borehole axis, means that it can
read-out of the two contrasting velocities, together with sense jointing, and stress-induced fracturing, that are
the energy anisotropy, the slowness anisotropy, and the missed by conventional logging tools. This means that
time anisotropy. Large energy differences, with the min- it is particularly useful for registering the additional
imum staying low, signal strong anisotropy. Both the jointing and fracturing that tends to be present on
acoustic time anisotropy and the slowness anisotropy either side of a fault. (Endo et al., 1997)
Rock stress, pore pressure, borehole stability and sonic logging 315

Figure 12.27 Dipole dispersion cross-over of fast and slow shear waves, demonstrated by modelling, and by testing unloaded and loaded
blocks of Berea sandstone containing a borehole. Caused by unequal radial stress gradients in the two principal stress direc-
tions, when uniaxially loading one of the blocks of sandstone containing an instrumented borehole. Winkler et al., 1998. This
study should ideally be performed on a borehole drilled while the block is under stress, e.g. following Addis et al., 1990,
Haimson, 2003.

A case was described from an Egyptian oil-producing with depth, the slowness data will tend to diverge from
well drilled in basement granites. With the shear-wave the expected trend (Brie et al., 1998).
anisotropy logging, the fault signature was clear, with By the beginning of the 21st century, it has become
the fast azimuth starting to change slowly when about more common to use dipole sonic logs to help interpret
20 m from the fault, with a rapid change of nearly 65° AVO or amplitude variation with offset anomalies, and
across the fault, followed by a return to an intermediate to help tie in ocean bottom seismic surveys (OBS) with
azimuth some 30 m beneath the fault, and a final return four-component (4C) acquisition. These techniques will
to background azimuth at greater distance. The particu- be briefly described with case examples, in Chapter 14.
lar fault had high permeability. The use of both wireline dipole logging in vertical
A final mention will be made of two other very use- holes, and pipe-conveyed dipole tools for deviated and
ful applications of LWD with dipole shear-wave horizontal wells has now given reservoir geophysicists
anisotropy analysis, which is available almost in real- improved means of calibrating the responses of their
time (some hours behind the drill-bit). The first is the rock physics based reservoir models, against small-scale
application to drilling of horizontal well sections, measurements.
designed to intersect a maximum amount of structure,
and simultaneously avoid the less favourable parallel to
H max hole direction. The second is early warning 12.7.1 Some further development of
while drilling in formations with rapidly changing pore logging tools
pressure. LWD is then an important aid in choosing
appropriate mud pressures. When porosities are no longer More recent descriptions of mechanical damage and
maintaining a normal trend of increased compaction anisotropy evaluation using a new modular dipole sonic
316 Rock quality, seismic velocity, attenuation and anisotropy

logging tool developed by Schlumberger, were given by


Pistre et al., 2002. This tool provided a complete seis-
mic characterization in radial, axial and azimuthal
directions. Improved monopole and cross-dipole tech-
nology, together with 13 axial levels, each with 8
azimuthal sensors, with each of the 104 receivers indi-
vidually digitised, resulted in 104 waveforms per trans-
mitter firing. Five transmitters are used. As the authors
emphasised, radial rock property variations arise
because of non-uniform stress distributions and
mechanical and chemical near-wellbore ‘alteration’ due
to the drilling process. The development of a discontin-
uum close to the wellbore is not mentioned, but per-
haps implied in the word ‘alteration’.
The authors emphasised that radial gradients of
acoustic slowness arose due to wave speeds being a
function of stress. (To this could be added the velocity
reducing effects of a discontinuous medium, due to rock
failure and various possible deformation mechanisms,
as illustrated earlier). The authors pointed out that
anisotropy could also be caused by intrinsic shale
anisotropy, possibly combined with external differen-
tial stresses. Inversion for these anisotropic, radial and
azimuthal rock properties could now be achieved from
the various acoustic behaviours, and from use of broad-
band dispersion curves.
A particular feature of the new tool was that profiling
of radial variations in compressional slowness could be
achieved through monopole acquisition with a wide
range of transmitter-receiver spacings, from very short Figure 12.28 Part of a dipole-generated flexural wave monitoring
log, here split into fast (left) and slow (right) shear
to very long. Radial inversion of shear slowness was
directions by stress and/or fracturing. Differences
quantified through inversion of the broadband disper-
are expressed here as slowness, for different radial
sions of the dipole flexural (Figure 12.25) and Stonely depths behind the wellbore wall. Pistre et al., 2002.
wave modes. When radial gradients were detected, in
e.g. sandstones, there could be an increased risk of
sanding, while lack of radial property gradients sug- radii from the wall of the borehole. The upper half of
gested mechanically intact rock, far from failure. The the profile shows a large radial gradient, from high
question of tangential-stress enhanced velocities and slowness near the wellbore, to a lower slowness at a radius
moduli, capable in principle of adding to the velocity of 24 in, or 63 cm. At the base of the profile the indi-
oscillation seen in Figure 12.11, was not mentioned. (As vidual curves tend to overlay, indicating a sounder rock
shown in Chapter 5, Figure 5.36, the modulus that is with negligible radial gradient of shear velocity.
influenced by cracking or jointing, shows strong stress-
sensitivity).
An example of part of a log display from an explo- 12.8 Mud filtrate invasion
ration/development well, is reproduced (in grey-scale)
in Figure 12.28, from part of a very comprehensive, The related themes discussed and illustrated in the for-
colourful log display given by Pistre et al., 2002. This going pages:
shows the result of the dipole flexural wave split into
fast (left) and slow (right) shear directions. The six-tone ● intensely fluctuating sonic log velocities in inter-
curves represent slowness (expressed as
s/ft) at different bedded (sandy?) shales
Rock stress, pore pressure, borehole stability and sonic logging 317

invaded zone would be important for processing and


interpretation of logs. The problem is caused by the
invaded (or ‘altered’) zone being deeper than that illu-
minated by the logging tool, meaning that the veloci-
ties will not reflect those of the formation, but of the
damaged zone, therefore requiring corrections.
They used multilayer velocity models to interpret
well measurements, for example the following layer
depths and velocities for the case of a slow and fast
formation:
Figure 12.29 Conceptual mud-filtrate invasion in a permeable
rock formation. After Chi et al., 2004.
Layer Radius Vp slow m/s Vp fast m/s

● mini-EDZ around experimental boreholes 1 0 1500 1500


● log-spiral shear surfaces observed in physical and 2 0.10 2300 4390
numerical models 3 0.18 2350 4512
4 0.26 2400 4634
● deformation (shearing, opening) of joints close to
5 0.34 2450 4756
numerically modelled wells
6 0.42 2500 4878
● damage zones confirmed by bicompressional
arrivals in real wells
These differed in a realistic manner from a more
suggest that mud-filtrate invasion modelling may per- commonly assumed sharp-interface model. In some
haps need to also consider flow through local discon- ways this small-scale gradational model mirrors the
tinua, such as interconnected log-spiral shear surfaces early controversy concerning stepped or gradational
in softer rocks, or less well- connected tensile and shear sub-ocean spreading-ridge velocity modelling, also
fractures in harder rocks, and flow along bedding applying to an unchanged rock type, in that case basalt,
planes and joints in general cases without over-stress. and applying to kilometre-scale depths. (Chapter 11)
Chi et al., 2004 mention the fairly common finding, Both phenomena are reflections of the changing
based on LWD and subsequent wireline logging differ- degree-of-fracturing, and of its interaction with the
ences, that near-wellbore formations are often altered by local effective stress.
stress, stress release and an assumed mud-filtrate invasion, Chi et al., 2002, emphasised that synthetic seismo-
as envisaged, schematically in Figure 12.29. However, a grams often did not correlate with measured seismo-
relatively uniformly-paced invasion, based on porosity- grams, when correlating seismic data with acoustic logs.
permeability conversion may perhaps be compromised It appeared that the standard approach here was to ‘cor-
by an actual permeability enhancement, based on the cen- rect’ the acoustic logs via a Biot-Gassmann fluid substi-
tral thesis of this book that rock quality, i.e. degree of tution, to ‘free sonic logs from mud-filtrate invasion
jointing, stress-induced fracturing, deformation mod- effects’. Doing this, it was assumed that the measured
uli, velocity and permeability, are quantitatively linked. velocities were those of the invaded zone, saturated
According to the authors, the ‘alterations’ cause the with mud filtrate. By ‘displacing the saturation fluid’
physical properties in the near-wellbore region to be theoretically, new velocities were obtained, and taken
different from those of the uninvaded rock formation. as the virgin formation velocities. However, there is a
In addition, stress concentration may cause formation potential problem here, if the near-wellbore is stress-
anisotropy, and an azimuthally varying radial variation fractured, despite ‘removal’ of the saturating fluid.
of velocities. As they point out, in well-consolidated In the example they described, surprisingly, only of
hard rock formations, mechanical damage is less pro- the order of 2 to 3% velocity changes resulted from the
nounced than in soft formations, so mud-filtrate inva- Biot-Gassman fluid substitution, showing that a reduc-
sion would then be more localized. tion of Vs and an increase for Vp had occurred in rela-
Formation properties inferred from wireline logging tion to true formation properties. This magnitude of
measurements may not reflect the true properties, so change seems immaterial in relation to the total effect
their opinion was that a realistic description of the of potential wellbore damage, clearly mostly caused by
318 Rock quality, seismic velocity, attenuation and anisotropy

‘rock mechanics’ effects (i.e. stress induced fracturing of same interval in the same well. Both monopole and
some form). dipole logs were measured by the wireline tool, which
In addressing the needs of a multi-physics approach was run some 10 days following completion of the
to the complex question of formation evaluation, based drilling. The LWD tool provided dipole measurements.
on time effects registered by well logging (LWD fol-
lowed by wireline), Torres-Verdin et al., 2003 utilised a
variety of modelling approaches. In Figure 12.30 they
show the result of a four-days mud-infiltrate invasion
model, using constant-permeability-with-radius assump-
tions, but with a 1:10 ratio of Kvertical and Khorizontal in
the formation of interest. The figure shows the calcu-
lated formation resistivity distribution, due to the unequal
rates of invasion in the central formation.
This anisotropy modelling gives an improved vision
of possible realities, but due to wellbore damage, mod-
ellers and analysts should perhaps also consider a
radius-dependent distribution of permeability, related
with the radius-dependent velocity caused by the (also
radius-dependent) degree of stress-induced fracturing
and stress redistribution. Such will only be necessary of
course, when the formation characteristics, in relation
to the stress levels and mud-pressure applied, cause
unwanted ‘alteration’ of properties, both immediately,
(a)
and exacerbated by time effects. Infiltration rates seem
likely to be accelerated at smaller radius, due to the pos-
sible presence of a miniature, failure-related discontin-
uum. (Revealing velocity-versus-radius measurements
in tunnels were reviewed in Chapter 7, see for instance
the circular tunnel cross-hole seismic results in Figure 7.5).
Briggs et al., 2004 in research at MIT, compared
LWD and wireline data that was collected over the

(b)

Figure 12.31 Scatter plots comparing higher frequency (shallower-


viewing, more disturbed) LWD tool velocities, with the
lower frequency (deeper-viewing, less disturbed) wire-
Figure 12.30 Modelling of four-days mud-infiltrate invasion, using line tool velocities. Briggs et al., 2004. In the case of the
constant-permeability-with-radius assumptions, but higher velocities, the earlier LWD measurement possi-
with a 1:10 ratio of Kvertical and Khorizontal in the for- bly views a less cracked and/or less invaded borehole
mation of interest. Torres-Verdin et al., 2003. wall, both tools registering relatively high velocities.
Rock stress, pore pressure, borehole stability and sonic logging 319

The authors found that for this well, there was on aver- The authors emphasised that in poorly consolidated
age about 5% discrepancy in shear wave data, and about zones that would be susceptible to damage by drilling,
3% in compressional wave data. These trends suggested the slower velocities should be accounted for, either by
higher average velocities from the subsequent, deeper- making a sufficient correction, or by using lower fre-
viewing wireline logs. A short section of the formation quencies and/or larger offset, in order to see deep enough
showed velocity differences of just over 10%, with an into the formation to register undisturbed velocities.
opposite trend. Significantly, both tools measure in dif- Material reviewed in this chapter has demonstrated
ferent frequency ranges, and had different offsets between that geomechanics/rock mechanics wellbore stability
source and first receiver. studies have an important place in improved understand-
The authors pointed out that as a general rule-of- ing of well-logging anomalies (e.g. Fjaer et al. 1992).
thumb, a tool sees 1 inch into the formation for every foot However, the existence of a miniature, but potentially up
separating the source and first receiver. Furthermore, low to several diameter EDZ discontinuum, which cannot be
frequencies (1–3 kHz) see 2 to 3 borehole diameters, modelled in conventional Mohr-Coulomb based contin-
while the higher frequencies see less than one borehole uum modelling (Barton, 2004b), due to incorrect addi-
diameter. Consequently, the higher frequency LWD tool tion of the cohesive and frictional components, is perhaps
was assumed to see the formation nearer to the borehole the root cause of the phrase ‘alteration zone’ being used to
wall than the lower frequency wireline tool. This would describe the complex, time-dependent interactions occur-
mean that the damage zone would tend to be seen by ring in the over-stressed, near-wellbore zone. The use of
the LWD tool, while the virgin formation would tend logging tools that illuminate to greater depth, due to low
to be seen by the lower frequency wireline tool. frequency, but that can also give information about the
However, the deeper penetration of split shear waves discontinuous zone, are clearly of importance for
nevertheless detects fracturing and anisotropy in the improved understanding of this cross-discipline region that
formation, outside the damage zone, as we saw earlier. surrounds wellbores.
The overall scatter plots of LWD versus wireline The mini-EDZ theme will be terminated with two
P- and S-wave velocities shown in Figure 12.31 support photographs from a large ‘borehole’, namely one of the
this radius-bias, with the wireline velocities sometimes world’s first TBM tunnels, excavated by a 7 ft (2.1 m)
1 km/s faster in Vp, and 0.5 to 1.0 km/s faster in Vs. diameter steam-driven machine credited to Beaumont,

Figure 12.32 One of the world’s first TBM tunnels, from 1880, credited to Beaumont. Excavation in chalk marl of UCS  4 to 9 MPa,
close to the Channel Tunnel between England and France. Note structurally controlled and (vertical) stress-controlled break-
out to at least one-radius on diametrically opposite sides of the tunnel. Barton and Warren, 1996.
320 Rock quality, seismic velocity, attenuation and anisotropy

for ‘pilot-drilling’ towards France, in an early (1880) Table 12.2 Definitions of maximum HPTP pressure and tem-
effort to examine the feasibility of a sub-sea link perature tiers (after Willson, 2006).
between England and France. Tier Reservoir pressure Reservoir temperature
The four pictures shown in Figure 12.32 are very
informative for this chapter. They show: a) structurally Tier I 15,000 psi (103 MPa) 350ºF (177ºC)
Tier II 20,000 psi (138 MPa) 400ºF (204ºC)
controlled, three joint set fall-out, which would be a
Tier III 30,000 psi (207 MPa) 500ºF (260ºC)
source of calliper-log ‘noise’ in other (vertical well) cir-
cumstances, disturbing the assumption of hole-ellipticity
parallel to h min. Photos b) and c) show progressive
The Tier II and Tier III categories are termed ultra-
stress-controlled ‘break-out’ at least doubling the effec-
HPHT and extreme HPHT for obvious reasons. The
tive diameter, caused in this case by tunnel loading under
current record for the offshore environment is report-
a 50–60 m chalk cliff along the S. England Folkestone
edly Mobile Bay, off the coast of Alabama, at 138 MPa
Warren coast line. The weak (UCS  4 to 9 MPa) chalk
and 215°C. There are now several North American deep
marl at tunnel level, has failed in combined shear and
gas reservoirs both onshore and offshore that would
compression, due to a vertical stress of only about 25%
classify as Tier III. As might be expected, there are oper-
of UCS, possibly with some bedding plane influence in
ating temperature limits for ‘conventional’ LWD log-
the ‘break-out’.
ging while drilling components, and for steerable
The discontinuum lies on the tunnel floor and behind
downhole drilling motors. Even Tier I temperature lim-
what we see at the tunnel wall. In a ‘mud-filtrate inva-
its may be reached at well depths in excess of 6 or 7 km,
sion’ scenario, the fluid would not penetrate in an
meaning that real-time data may no longer be available
isotropic or linear manner, but faster into the haunches,
while drilling. However, subsequent wireline logging
or ‘E-W’. In a further possible illustration of unwanted
can presently be performed at temperatures up to the
hole collapse, this time due to ‘over-pressure’, photo-
Tier III ‘limit’ of 260°C.
graph d) shows failure due to successive bedding plane
As we shall see in Chapter 13 on rock physics phenom-
opening under the sea section of this tunnel. The
ena at extreme levels of confining stress, the high state of
original circular 2.1 m diameter hole has ‘migrated’
compaction at great depth means that the porosity and
with time, into the collapsing, bedded rock above the
seismic velocity of typical reservoir rocks may exhibit
crown. Could this be the occasional fate above and
little sensitivity to changing depth. The prediction of
below horizontal sections of wells waiting for casing
fluid pressures and fracture gradients (e.g. Figure 12.1) is
installation, in case of thinly bedded strata with over-
then more difficult.
pressure?
A related problem of extreme rock pressure is that the
Terzaghi, 1943 theory of effective stress:

  – p (12.1)
12.9 Challenges from ultra HPHT
may no longer give a correct description of the magni-
The drive to discover more petroleum, which is especially tude of the three effective principal rock stress compo-
relevant at the time of publication (2006) with oil prices nents that would normally define the likely stability or
above 70 US $ per barrel, is stimulating the exploration of instability of a deep well. At great depth, the rock skele-
deeper petroleum reserves, with all the associated difficul- ton may bear a greater proportion of the total rock
ties of high temperatures, high fluid pressures, and high stresses due to an effective stress parameter  that is less
rock stresses. Some of this is so-called ‘infrastructure-led than 1.0, following the Biot, 1956b generalized theory
exploration’, searching deeper or laterally and deeper from of poroelasticity:
developed fields with their existing production facilities.
Since a 1990 accident report (the Piper Alpha plat-   – p (12.2)
form disaster in the UK sector of the North Sea), there
have existed formal definitions of high pressure high A ‘moderate’ example from Hettema and de Pater,
temperature (HPHT) wells. With time this has been 1998 is a clay-rich sandstone with an unstressed porosity
extended to the following three-tier system for classify- of 20%, which demonstrated an -value of 0.9 under
ing these extreme conditions. zero differential pressure, reducing to 0.6 at 36 MPa
Rock stress, pore pressure, borehole stability and sonic logging 321

differential pressure. A more extreme example is a lime- as discussed by Pepin et al., 2004 and others, in recent
stone tested by Laurent et al., 1993, with -values reduc- literature.
ing from 0.8 to 0.2 as porosity decreased from 23% to As pointed out by Willson, 2006 in a helpful techni-
4.5%. Such effects may increase the effective stress com- cal review of the subject (‘Feeling the heat, can’t stand
ponents by tens of MPa, which leads, as also in TBM tun- the pressure?’), manipulating the mud temperature dur-
nelling (Barton, 2000) to reduced drilling penetration ing the drilling process, before formation temperature is
rates as the rock is stronger. However it also leads to the re-established, can also be used ‘in the opposite direc-
opposite effect of a potentially increased likelihood of tion’ to positively influence the hydraulic fracture (mud-
stress-induced fracturing or (log-spiral) shearing, depend- loss) gradient if there is no risk of formation well-bore
ing on rock type. (In the world of TBM tunnelling there compressive failure. This would be done by increasing the
are extremes of 0.1 m/hr and 10 m/hr penetration rates, mud temperature. However, in the case of shale, an ini-
the former due to extremely hard rock at great depth in an tially undrained condition when exposed by drilling
Idaho mine, with a TBM giving insufficient thrust. means that the greater thermal expansion coefficient of
Practical measures for reducing the double impact of the contained water will have greater influence on the
both high effective stresses, and the additional thermo- stability than the rock’s response to increased or lowered
elastic effect of high drilling mud temperatures were pro- mud temperatures (Li et al., 1998).
posed by Maury and Guenot, 1995, by introducing mud There are reportedly particular problems when pene-
cooling systems. When the circulating mud is cooler trating salt rocks at high temperature and pressures
than the formation, thermoelastic contraction means (Willson and Fredrich, 2005), due to the particular
lower tangential stresses, with the dual effect of reducing sensitivity of the creep rates of salt to high temperature.
the likelihood of compressive stress-induced fracturing, There is a so-called ‘undefined mechanism’ of creep at
but an increased likelihood of mud pressure induced ten- lower temperatures, and a ‘dislocation climb mecha-
sile cracking, due to the reduced minimum tangential nism’ at high temperatures, which can result in orders of
stresses on opposite sides of the well, and rotated 90° magnitude increased rates of ‘creep’. Already between
from the potential compressive fracturing locations. 60°C and 230°C there is a reported 200 increase.
The tensile cracking allows mud loss, but some is Clearly this is of particular concern to petroleum com-
returned when the temperature subsequently rises. The panies who are developing reserves beneath thick bed-
apparent wellbore ‘ballooning’ represents the mud loss, ded salt formations or next to salt dome structures. A
prior to the subsequent gain of fluid. The subject of review of drilling problems, and an emphasis of the need
mud temperature management is of great current inter- for high pressure and high temperature creep tests for
est for ultra deep wells, for extending fracture gradients, salt is given by Maia et al., 2005.
13 Rock physics at laboratory
scale

King, 2005 recently summed up the major exploration- as velocity-versus-porosity, porosity-versus-permeability,


related goals of rock physics research. They are ‘to under- Vp/Vs-versus-saturation and lithology. The author dis-
stand how lithology, porosity, confining stress and pore cussed the seismic signatures of cementation, sorting,
pressure, pore fluid type and saturation, anisotropy and shaliness, lithology, fluid content, and compaction.
degree of fracturing, temperature, and frequency influ- Both the well-established porous matrix behaviours,
ence the velocities and attenuation of compressional and several of the newer areas of knowledge such as joint
P- and S-waves in sedimentary rocks’. At the end of this and fracture behaviour, will be reviewed in this chapter,
list the author added ‘and vice versa’, presumably empha- before moving to reservoir-scale in Chapters 14 and 15.
sising the interactive and complex nature of the reality. Of necessity the selection of material is subjective, but
Isolation of just two or three of these variables for intense, designed to be informative, and broad-based, within the
high quality investigation, clearly gives just a small incre- confines of one chapter. A lot of supplementary data will
ment to the overall understanding. be found concerning seismic Q, to add to the introduc-
His review article ended by emphasising the remaining tory material of Chapter 10. In this chapter it is presented
challenges: relationships between attenuation, anisotropy, together with the velocity data.
fractures and fluid flow – and determining these relation- A ‘shallow perspective’ of this subject, certainly in rela-
ships across the frequency spectrum of core, log and seis- tion to depth of observation, will be found in Chapter 2,
mic measurements. In this chapter we will review some of based mostly on civil engineering related investigations.
the important increments in this understanding, mostly by A certain relation to many of the high-pressure velocity-
first treating velocity-related experiments, then attenu- parameter trends presented in the present chapter will be
ation-related experiments. Description of various matrix noted. The reader is referred to some particularly interest-
behaviours will be followed by description of laboratory ing assemblies of low-stress data. The particular effects of
tests that include jointing or induced fracturing. variable weathering and fissuring common to the near-
There was early recognition (e.g. Birch and Bancroft, surface is documented, which is absent in reservoir-related
1938), that seismic velocities of rocks were strongly influ- studies, with the possible exception of related effects from
enced by microcracks, and that seismic attributes repre- alteration close to fault zones, that may exhibit hydrother-
sentative of the intrinsic mineralogy and porosity, could mal alteration of their various mineral assemblages.
only be obtained by applying pressure to the rocks. Much
of the rock physics understanding of reservoir rock behav-
13.1 Compressional velocity and
iour (of both matrix and joints), has therefore to be
porosity
achieved at elevated pressure, and the importance of
temperature is also well recognised, but less frequently
In the petroleum industry there is a history of at least
an experimental variable.
half a century for inferring porosity from well logs, and
Mavko, 2002 pointed out that because of the grow-
for indicating pore fluid type. Among the oldest, and
ing complexity of recently discovered oil fields, a major
most popular expressions was that put forward by Wyllie
shift had taken place in the use of seismic methods in
et al., 1956, based on time-averaging in the solid and
the last decade of the 20th century. Interpretation of the
fluid phases. Using Tp to represent the total travel time of
increased spatial variability due to heterogeneous distri-
a P-wave, and Ts and Tf to represent the travel time com-
butions of porosity, clay content, fracture density and
ponents expected in the solid and fluid phases respectively,
permeability, was now more strongly based on rock
it is assumed that:
physics understanding. Gradually, more order had been
discovered in relations that once seemed scattered, such Tp  Ts  Tf (13.1)
324 Rock quality, seismic velocity, attenuation and anisotropy

A ‘reservoir rock’ selection of P-wave velocity versus


porosity data for saturated samples, is reproduced in
Figure 13.2 for ready reference. This limited Vp- selec-
tion, from Mavko et al., 1998 appendices, is also credited
to original authors in the figure caption. Chalk, two sets
of limestone data of widely different porosity ranges,
dolomite, and two sets of sandstone data are reproduced.
As we shall see shortly, explaining some of the wide scat-
ter in the data requires specification, as a minimum, of
effective stress level, degree of cementation, and clay-
content. Examples of such discrimination are shown in
Figure 13.3, from Dvorkin and Nur, 1998.

13.2 Density, Vs and Vp


Figure 13.1 P-wave velocity versus porosity for compact clean sand-
stones, and for a suspension of component grains, The age-depth relationships derived for numerous wells
behaviour that is separated by the critical porosity c in sandstone-shale units by Faust, 1951 (see beginning
concept. Nur et al., 1991, and Mavko et al., 1998. of Chapter 14), had a certain grouping of velocities with
age, due to variations of porosity and the resulting dens-
Inserting the fractional porosity () of the rock, and Vp ities. Hard porosity in the form of pores tended to
for measured velocity, the component velocities Vps and decrease with age and depth, while (soft) porosity in the
Vpf contribute as follows by substitution in equation 13.1: form of joints tended to increase with age due to tec-
tonic influences, but reduced strongly with depth. Only
1 1   the hard porosity will have a significant effect on density.
  (13.2)
Vp Vps Vpf Gardner et al., 1974, showed simple velocity-density
trends for reservoir rocks, which on a log-log plot were
A close variant of this Wyllie et al., 1956 time-average almost linearly distributed. Figure 13.4 shows the trends
prediction, is the popular velocity-to-porosity equation for sandstone/shale/limestone/dolomite, where the cen-
of Raymer et al., 1980, for   0.37: tral relationship (dashed line), as in equation 13.4 is:

Vp  (1  )2 Vps   Vpf (13.3) g  0.23V


1
4 (13.4)
However, as pointed out recently by Dvorkin and Nur,
1998, in an appropriately titled note: ‘Time-average equa- where   bulk density in gm/cm3, and velocity Vp is ft/s.
tion revisited’, there is actually only a limited theoretical (Note: density  velocity  impedence  gm/cc  m/s)
scope for such time-average equations, requiring arrange- We can visualise more familiar (km/s) velocities and
ment of the two components in layers normal to the densities by estimating  when Vp is 3, 4 or 5 km/s
direction of propagation, with a wave length small com- (9,840, 13,120 or 16,400 ft/s). The three results for
pared to the thickness of an individual layer.  according to this equation are approximately 2.29,
Nur et al., 1991 championed the concept of a criti- 2.46 and 2.60 gm/cm3 respectively. When Vp  6 km/s
cal porosity that separates both the mechanical and a density of 2.72 gm/cm3 is predicted. Each of the
acoustic state of a rock and its component grains. At above is a realistic reflection of both mineralogy and
lower porosities than c the mineral grains are load- porosity differences, as density and velocity increase.
bearing, giving correspondingly higher velocities, while Several sets of seismic data that included the funda-
at porosities greater than c a fluid phase will be load- mental effect of density were presented in Chapter 2,
bearing, the particles being in suspension, with the obvi- but these were not specifically related to reservoir rock,
ous consequences of a very ‘flat’ velocity-porosity and were usually at ‘near-surface’ stress levels, related to
response. An example of this concept, for the case of civil engineering projects, and often had the density- and
clean sandstones, is given in Figure 13.1. velocity-reducing influences of weathering.
Rock physics at laboratory scale 325

Figure 13.2 An assembly of Vp-porosity data for saturated chalk, two limestones of widely different porosity, dolomite, and two sandstones
of medium and high porosity. a) Chalk: from sonic log and porosity log, Urmos and Williams, 1993. b), c) Limestones: ultrasonic,
10 to 50 MPa effective pressures, Cadoret, 1993, Lucet, 1989, Yale and Jamieson, 1994. d) High porosity sandstone: ultrasonic,
35–40 MPa effective pressures, Strandenes, 1991. e) Sandstones (with yet-to-be-defined clay-content): ultrasonic, 30–40 MPa
effective pressures, Han, 1986. f ) Dolomites: ultrasonic, 10–35 MPa effective pressures, Geertsma, 1961, Yale and Jamieson,
1994. After Mavko et al., 1998.

A ‘reservoir rock’ selection of P-wave velocity versus mineralogy of ‘sandstones’, with 10–15% variation in
density data, is reproduced in Figure 13.5a for ready density possible for the same velocity, particularly in
reference. This limited Vp- selection, from Mavko et al., the case of the tight gas sandstones, which gives a cor-
1998 appendices, is also credited to original authors in relation coefficient of only 0.39. In contrast to these
the figure caption. Chalk, limestone, dolomite, and three variations, the Vs  Vp trends are consistently uniform,
sets of sandstone data are reproduced. Again for refer- as befits characterization by seismic waves.
ence purposes, both the shear-wave and compression- In Chapter 14, an early introduction to the age
wave velocities, for the same six groups of rocks, are and depth effects on velocity is reproduced from Faust
shown in Figure 13.5b. 1951, who analysed well survey results from some 500
The relatively ‘ordered’ density-Vp trends for the chalk, petroleum wells in the USA and Canada. Faust used
limestone and dolomite reflect the simpler mineralogy. data from about 300 kilometres of well sections. The
The contrast to the widely scattered density-Vp data for great majority of data was for mixed shale/sandstone
the three groups of sandstones is evidence of the variable sections.
326 Rock quality, seismic velocity, attenuation and anisotropy

A non-systematic comparison of shale and sand (sand-


stone) velocities revealed an average discrepancy of only
350 ft/sec, or 106.7 m/s in velocity between these two,
frequently inter-bedded units, the sandstone having the
highest velocity by this small average margin. We will
present the also remarkably close Vp versus Vs trends for
water-saturated sandstones and shales, from Castagna et
al., 1993, as reproduced in Mavko et al., 1998, at the
appropriate location in Chapter 14 (Figure 14.4), to
emphasise the remarkably similar Vp and Vs signatures of
these two ‘dissimilar’ lithologies, when in a compacted
state. The necessity of using impedence (gm/cc  m/s),
attenuation, and anisotropy, for seismically distinguishing
these two most essential reservoir ‘partners’ is clear.

13.3 Velocity, aspect ratio,


pressure, brine and gas

An important early paper in the area of theoretical


modelling of porous rock behaviour was presented
Figure 13.3 Specification of sandstone condition: quartz-cemented,
by Toksöz et al., 1976, who examined the numerous
clay-cemented, uncemented, and specification of effect-
ive stress level (applied on uncemented Troll sand),
factors affecting seismic velocities of intact samples of
helps to sort P-wave velocity data that displays ‘unex- porous rocks with emphasis on sandstones. They devel-
plained’ scatter. Dvorkin and Nur, 1998. The sub- oped theoretical formulations to represent the solid
envelopes beneath the Troll sand data are from Dvorkin matrix, and assumed spherical to oblate spheroidal pores,
and Nur, 1996. of widely varying aspect ratios, to match numerous labor-
atory data. As one would expect, they found that small
aspect ratios (flatter voids) caused greatest reductions to
elastic moduli and velocities.
They also predicted and confirmed that the properties
of the saturating fluid (gas, oil or water) produced greater
effects on the compressional velocities than on the shear
velocities. The P-wave velocities were predicted, correctly,
to be higher when the rock was saturated with water,
than when dry or gas-saturated.
When fitting their theoretical model to P- and S-wave
velocities that were measured at different pressures,
they required pore shape spectra ranging from spheres
to very fine cracks (aspect ratios from 1 to 105) for
sandstones, limestones and granites, both under dry
and saturated states. As igneous rocks have low porosi-
ties, the pore shape has great influence on the elastic
and seismic properties, and dry and water-saturated
behaviours are markedly different, as was also seen in
Chapter 2.
Compressional velocities were highest with brine sat-
uration and lowest with gas saturation, but the differ-
ence declined with increasing pressure. Poisson’s ratios
Figure 13.4 Log-log trends for Vp (ft/s), and . Gardner et al., 1974. for gas saturated rocks were lower than for those with
Rock physics at laboratory scale 327

(a)

Figure 13.5a An assembly of Vp-density data for saturated chalk, limestone, dolomite, and three sandstones of low, medium and high poros-
ity. a) Chalk: from sonic log and porosity log, Urmos and Williams, 1993. b) Limestone: ultrasonic, 10 to 50 MPa effective pres-
sures, Cadoret, 1993, Lucet, 1989, Yale and Jamieson, 1994. c) Dolomites: ultrasonic, 10–35 MPa effective pressures, Geertsma,
1961, Yale and Jamieson, 1994. d) Tight gas sandstones: ultrasonic, effective pressures 40 MPa, Jizba, 1991. e) Sandstones:
ultrasonic, 30–40 MPa effective pressures, Han, 1986. f ) After Mavko et al., 1998. f ) High porosity sandstone: ultrasonic,
35–40 MPa effective pressures, Strandenes, 1991. After Mavko et al., 1998.

brine saturation, and this difference persisted to great 50 MPa. Greatest sensitivity, as one would expect, was
depths according to their model. shown when pressure or depth was smaller.
Figure 13.6, from Toksöz et al., 1976, is a good exam- From summaries of the numerous experimental and
ple of their modelling predictions, showing the relative theoretical trends given by Toksöz et al., 1976, Table 3.1
predicted effects on Vp of brine-filled and gas-filled cracks was developed.
in a 16% porosity sandstone model. The fluid occupy- The effects of the dry or brine-saturated states, and
ing the smaller aspect ratio cracks has more influence the influence of effective stresses, as predicted theoret-
on velocities at low pressures, due to the greater pres- ically by Toksöz et al., 1976 so long ago, were nicely
sure sensitivity of the fine cracks. illustrated by more recent testing by King and Marsden,
Figure 13.7 shows the same authors’ predicted Poisson’s 2002, who tested numerous sandstones both dry and
ratio variations, with varying degrees of brine or gas sat- brine saturated. Ultrasonic P- and S-wave measure-
uration, and also as a function of differential pressure to ments were made on ten sandstones with porosities less
328 Rock quality, seismic velocity, attenuation and anisotropy

(b)

Figure 13.5b An assembly of Vs - Vp data for the same saturated chalk, limestone, dolomite, and three sandstones of low, medium and
high porosity, that are presented in Figure 13.5a. a) Chalk: from sonic log and porosity log, Urmos and Williams, 1993.
b) Limestone: ultrasonic, 10 to 50 MPa effective pressures, Cadoret, 1993, Lucet, 1989, Yale and Jamieson, 1994. c) Dolomites:
ultrasonic, 10–35 MPa effective pressures, Geertsma, 1961, Yale and Jamieson, 1994. d) Tight gas sandstones: ultrasonic,
effective pressures 40 MPa, Jizba, 1991. e) Sandstones: ultrasonic, 30–40 MPa effective pressures, Han, 1986. f ) After Mavko
et al., 1998. f ) High porosity sandstone: ultrasonic, 35–40 MPa effective pressures, Strandenes, 1991. After Mavko et al., 1998.

than 10%, and thirty-four specimens with porosities in state, given by the authors, has been omitted from these
the range 20 to 30%, under hydrostatic effective stresses figures for the sake of clarity.
up to 60 MPa (in the case of the stronger, lower poros-
ity samples), and up to 40 MPa in the case of the higher
porosity set. 13.4 Velocity, temperature and
Figure 13.8 shows the Vp and Vs results for the dry and influence of fluid
saturated specimens to a common hydrostatic stress of
40 MPa. Equations relating Vp and Vs are shown in the Although far from exhaustive, a limited set of data for
figure. temperature effects on hydrocarbon-saturated samples
King and Marsden, 2002, also presented their Vp and will now be presented to illustrate some of the geophys-
Vs results as a function of the effective stress (10, 20, 40 ical changes that can be used to monitor producing, stim-
or 60 MPa). Figure 13.9 shows both the Vp-Vs-effective ulated reservoirs. Nur, 1989, referred to the new ‘four
stress trends, and the dry-saturated trends. The add- dimensional seismology’, in other words the ability to
ition of a ‘Gassman predicted’ result for the saturated monitor in three dimensions the effect of time during
Rock physics at laboratory scale 329

Table 13.1 Typical relative effects of environmental conditions for


porous reservoir rocks (derived from Toksöz et al., 1976).

Lower velocity (Vp) if Higher velocity (Vp) if

Low water saturation High water saturation


Dry or gas saturated Dry or gas saturated
(if flatter pores) (if rounder pores)
Some immiscible gas (in brine) Saturated with brine
Higher porosity No immiscible gas
Over-pressured Lower porosity
Shallow depth Under-pressured
Thin pores Greater depth
After several cycles of freezing Rounded pores
Room temperature Frozen
Extremely high temperature Low or moderate temperature
Lower velocity (Vs) Higher velocity (Vs)
Figure 13.6 Theoretical variations of Vp with differential pressure
If water saturated If dry or gas saturated (and
(0 to 50 MPa) and different aspect ratio cracks, for var-
highly porous)
ious brine or gas saturation levels. Toksöz et al., 1976.

Figure 13.7 Toksöz et al., 1976 model predictions for Poisson’s


ratio as a function of pressure or depth, and as a func-
tion of the degree of saturation with brine, gas or
a mixture of the two.

various flooding methods. The basis for such an ability


would be the strong dependence of velocity on tempera-
ture, (water-flooding causing local cooling, steam
flooding the opposite), plus the significant influence of
the relative hydrocarbon and brine saturations.
The six sets of results shown in Figure 13.10 show
an easily detectable effect of temperature, with greatest Figure 13.8 Vp and Vs as a function of condition (dry or brine-
effect when 100% oil saturation, and least effect (almost saturated), at a common hydrostatic effective stress of
zero effect) when 100% gas saturated or 100% brine 40 MPa. King and Marsden, 2002.
330 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.10 Vp behaviour as a function of relative hydrocarbon


Figure 13.9 Vp and Vs as a function of effective hydrostatic stress, saturation and temperatures up to 150°C. Top: Kern
for a) ten lower porosity sandstones, b) thirty four high River oil sand: P  100 bars, PP  0 bars, Bottom:
porosity sandstones. Dry and brine-saturated results Venezuelan oil sand: P  100 bars, PP  30 bars.
are shown. Redrawn from King and Marsden, 2002. Nur, 1989.
Rock physics at laboratory scale 331

saturated. This temperature effect on velocity reduces dependence of the viscosity of the oil. The phenomenon
to about half strength when 50% oil/50% gas or 50% was described as follows by Winkler and Nur. At low
oil/50% brine are present. In other words, when the oil temperature the higher viscosity means that the oil can-
is removed from these sands, the velocities successively not flow easily, so the dynamic measurement is on the
become independent of temperature, with roughly half high-frequency, high velocity, unrelaxed side of the local-
the effect when 50% oil remains. flow mechanism. As temperature increases viscosity
Nur, 1989, also referred to a pilot steam flood experi- reduces, so fluid flows more easily, and velocity there-
ment in viscous tar sands in Canada, conducted by fore decreases since measurement is on the relaxed side
Amoco. Changes in travel-time caused by heating effects of the absorption/dispersion mechanism.
reducing Vp were readily detected. A differential travel-
time plot revealed the areas closest to the wells (and
13.5 Velocity, clay content and
partly between the wells) where velocity had been
permeability
reduced by the change in viscosity. Laboratory tests of
the tar sand showed an S-shaped reduction in velocity
A comprehensive series of laboratory tests reported
from Vp  2.4 km/s at 25°C, to a final plateau of about
by Klimentos, 1991, were designed to investigate the
1.1 km/s at 150°C. (Den Boer and Matthews, 1988).
influence of clay content on the P-wave velocities of sat-
In their review of acoustic velocity and attenuation in
urated sandstones under varying confining and pore fluid
porous rocks, Winkler and Nur, 1995 refer to the work
pressures up to 40 MPa. Forty-two samples of sandstones
of Wang and Nur, 1990 who measured the temperature
were investigated, having the following range of charac-
dependence of compressional and shear wave velocities
teristics, in order to see their combined and individual
in sandstones that were either dry, saturated with water,
effects on velocity:
or saturated with crude oil. Figure 13.11 shows the quite
widely differentiated results, due again to the temperature ● Porosity: 2 to 36%
● Permeability: 0.001 to 306 mD
● Clay content: negligible to 30%
The principal results are shown in Figure 13.12, with
sorting according to clay-content, shown in Figure 13.13.

Figure 13.11 Compressional and shear wave velocities in Boise


sandstone, as a function of temperature and saturat-
ing fluid. The samples were dry, or saturated with
water or saturated with crude oil. A common effective Figure 13.12 P-wave velocity at 1 MHz frequency and 40 MPa
stress of 15 MPa was applied. Wang and Nur, 1990 confining pressure, showing clay contents and
in Winkler and Murphy, 1995. porosities. Klimentos, 1991.
332 Rock quality, seismic velocity, attenuation and anisotropy

where   porosity (fraction), c  clay content (frac-


tion) and K is permeability in millidarcies.
The effect of permeability alone on the P-wave vel-
ocity is seen to be negligible. In fact a misleading, weak
increase in Vp with increasing permeability arises due to
the velocity-clay content and clay content-permeability
interrelations. Clearly these equations can be reformu-
lated to give estimates of (matrix) permeability.
Klimentos and McCann, 1990, also drew attention to
the complex nature of permeability – depending as it did
on porosity, pore size distribution, inter-connectedness
of the pores, and tortuosity – the latter two presumably
being especially compromised by clay content. They
also posed as an open question, what the relative effects
(a)
would be of frequencies of 10 Hz to 1 kHz (as used in
seismic exploration) or frequencies of 10 to 20 kHz (as
used in borehole logging), when clays were present in
the in situ sandstones.

13.6 Stratigraphy based velocity to


permeability estimation

Gutierrez et al., 2002, also addressed the question of


clay-content, referring to the initially undifferentiated
data set of Han, 1986 for sandstones. Figure 13.14 shows
the undifferentiated Vp versus porosity data of Han,
1986: top-left, and the stratigraphy-guided, clay-content
differentiation: top-right, that makes the nearly 2 km/s
variation in Vp at one porosity, understandable.
(b) Jan, 2003 presented a finer clay-content break-down
of Han’s data, including some additional data, which is
Figure 13.13 a) P-wave velocity as a function of clay content for included in Figure 13.14c and d. Both the P-wave and
porosities of 6 to 36%. b) P-wave velocity as a func- S-wave velocities were measured at 40 MPa confining
tion of clay content with average 15% porosity pressure and 1 MPa pore pressure.
(squares) and 28% porosity (circles), each at 1 MHz
In order to emphasise the potential of some stratigraphy-
and 40 MPa confining pressure. Klimentos, 1991.
guided lab-to-field velocity-porosity-permeability corre-
lations, Gutierrez et al., 2002 presented a well log for
From Figure 13.13, where porosity differences were
La Cira-Infantas Oil Field (LCI). This is reproduced
used to distinguish the effect of clay content. Klimentos,
in Figure 13.15. The left-hand velocity log applied to the
1991, gave the following multivariable linear regression
highly variable (also laterally) clean-to-shaly, loosely con-
equations for estimating Vp:
solidated, Tertiary sandstones, lying in a highly faulted,
asymmetrical anticline. Well-logged Vp ranged mostly
1. at ultrasonic frequencies (1 MHz):
from 2.5 to 4 km/s, while core-based porosity was mostly
between 10 and 30%, and there was a four order-of-
(r  0.96) Vp  5.66  6.11  3.53c
magnitude range of core-based permeability. The core-
 0.0007K (13.5)
based results are plotted in Figure 13.15b and c.
2. at seismic frequencies:
The authors plotted a series of velocity-porosity dia-
(r  0.93) Vp  5.27  5.40  2.54c grams, starting with an 850 m interval with a few thou-
 0.001K (13.6) sand data points for the undifferentiated deposition
Rock physics at laboratory scale 333

(a) (b)

(c) (d)

Figure 13.14 a) Unsorted Vp – n% data for sands with 35% clay. b) Logical differentiation of Vp – n% trends when grouped by clay-content.
Gutierrez et al., 2002. c) and d) A more detailed presentation of Han, 1986, data given by Yan, 2003.

cycles of the Tertiary basin. Successively smaller depos- permeability, through a more relevant match of porosity
ition cycles were then considered, first down to a specific and permeability with common sediment compaction
operational zone, then so-called ‘fining-up’, giving suc- and cementation history.
cessively higher Vp-n% correlation coefficients, reflecting This work followed the permeability-porosity match-
the more uniform sedimentary environment and diage- ing of Amaefule et al., 1993, who showed the importance
netic nature of the smaller cycles. An essentially of separation into hydraulic units. Possibly we can draw
linear plot was shown for one of the fining-up cycles: in a parallel here to the common separation of rock mass
a 5 m section Vp was 3.5 km/s at n  10%, and 2.5 km/s qualities into classes (Q  1  4: poor, Q  4  10:
at n  30%, with correlation coefficient r  0.973. fair, etc), for different structural domains, which is the
The authors found that due to a large fraction of silt basis for prediction of similar behaviour, such as particu-
(whose mineralogy is close to that of clean sand), the lar reinforcement needs in a tunnel, a particular range of
clay-content in 100% shale intervals could be as low as deformation moduli or P-wave velocities.
20–30%. Poor sorting in the shale caused a reduction Amaefule et al., 1993 and Prasad, 2003 used the fol-
in total porosity, which caused an increase in the vel- lowing simple ‘classification’ relations:
ocity in relation to existing models, and in relation to the
data sets in Figure 13.14, which show lower velocities RQI  0.0314 (k/)
1
2 (13.7)
with these amounts of clay.
The concept of sorting data into common categories, 1
using stratigraphy and other matching techniques, was FZI  0.0314/ (k/) 2 (13.8)
also the theme in Prasad, 2003, who showed that by
grouping and sorting rocks into hydraulic units, it was RQI is known as the reservoir quality index, with
easier to establish relationships between velocity and permeability (k) in units of millidarcies, and () is the
334 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.16 Porosity-permeability correlation from core of a sec-


ond well in the LCI Tertiary sands, which allowed
sonic-logging data to be used for permeability pre-
diction. Gutierrez et al., 2002.

13.6.1 Correlation to field


processes

Figure 13.15 a) A well log from the heterogeneous Tertiary sands Prasad, 2003 also provided a practical illustration of the
of the LCI field, showing Vp, with stratigraphy- effect of depth-of-sediment on porosity, velocity and per-
guided core data for b) porosity and c) permeability. meability development to 500 m below sea floor, using
Gutierrez et al., 2002. marine logging results from ‘Site 977, ODP Leg 161
(Shipboard Scientific Party, 1996). In the parallel dia-
grams reproduced in Figure 13.19, the scattered down-
fractional porosity. The term () is the void ratio, given hole log data is shown beneath the smoothed trend lines
by /(1-). It thus links FZI the flow zone indicator A-B in each case. Curves D-D' represent the hypotheti-
with RQI, using the ratio of pore volume to solid vol- cal trend if porosity is ‘frozen’ by a cementation episode
ume. Rocks with FZI values within a narrow range from point D. Note the FZI-predicted maintenance of
belonged to one hydraulic unit: they had similar flow high permeability, and the ‘non-correlating’ increase in
properties. A semi-log plot of porosity versus perme- velocity (but one that can be explained by FZI).
ability showed similar FZI values plotting together. Curves C-C' represent the hypothetical effects of an
Prasad first tested this older method, showing ‘opposite’ trend – an influx of pore-filling materials from
unsorted data (left), then data sorted by FZI (right) point C. There is a rapid reduction in porosity. This can
in Figure 13.17. This demonstrated that the FZI con- be used to give an estimated reduction in velocity fol-
cept could be extended to seismic parameters, giving lowed by an FZI-predicted reduction in permeability.
a strong correlation between velocity and permeabil- Prasad, 2003 calculated the permeability scatter-curves
ity, when using the appropriate FZI grouping. She used directly from assumed FZI values of 0.1 (left curve),
a laboratory-test data base, which included porosity, and 0.25 (right curve), using the log data for porosity
permeability, velocity and attenuation data from tests and velocity. Note the similar scatter-shape of porosity
at similar confining pressures. Figure 13.18 shows data, and FZI-calculated permeability, due presumably
a much larger set of data with Vp – k correlation to the use of (porosity)0.5 and (void ratio)1.0 in the
through FZI. FZI estimate.
Rock physics at laboratory scale 335

(a) (b)

(c) (d)

Figure 13.17 a) Unsorted log k versus ␸ data on the left, and b) sorted data using the expression for FZI with measured matrix parameters
on the right c) Unsorted log k versus Vp data on the left, and d) sorted data using FZI on the right. Prasad, 2003.

13.7 Velocity with patchy saturation


effects in mixed units

Knight et al., 1998, showed that uniform or smooth vari-


ations of velocity with degree of saturation were strictly a
function of an assumed or actual homogeneous distribu-
tion of saturation due to lithological uniformity. They
investigated the more complex (and common) effects of
having mixed lithological units, which tended to create a
heterogeneous or patchy, saturation distribution, with
different signatures during imbibition and drainage.
With a more complex distribution of saturations due to
lithology differences, it was only when close to saturation
of 100%, that there was a consistent steep rise in velocity.
(From another field, and for harder rock types, see Saito,
Figure 13.18 Permeability estimated from velocity data, using 1981, in Figure 2.17a, b in Chapter 2 of this book).
FZI values from a larger data base of reservoir sand- The authors found that pore-scale and sample-scale
stones, marine reservoir sand and a tight sandstone. fluid distribution effects, and of course capillary effects,
Prasad, 2003. caused different Vp response (in degree but not general
336 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.19 An illustration by Prasad, 2003 of the possibility of predicting permeability from log data of porosity and permeability, using
a relevant, logging-based FZI-value. Marine logging results from ‘Site 977, ODP Leg 161 (Shipboard Scientific Party, 1996).
Hypothetical mineralization from D (curve D-D’) , and hypothetical pore-filling from C (curve C-C').

style) when draining as compared to imbibing. The allow reinforcement of the compliant part of the rock,
drainage process creates a more heterogeneous distribu- so velocities are low.
tion of saturation. At high frequencies, (Figure 13.20b), this pressure
Local full saturation of the crack-like regions of the equilibrium cannot occur because the pore fluid relax-
pore space tend to stiffen these regions in relation to high ation time is greater than the seismic wave period. Pore
frequency, but at low frequency these ‘patches’ can drain fluid in the thin compliant pores is then effectively
to the less saturated pore space. The phenomenon appears ‘trapped’, and it therefore reinforces the otherwise com-
to be shifted in frequency at ‘macroscopic-patch’ scale. pliant pore spaces, resulting in higher apparent modu-
Such results clearly impact poroelastic modelling with lus and velocity.
different frequencies, which is discussed in Chapter 15. Knight et al., 1998 described a comprehensive inves-
Knight et al., 1998, found that the pore fluid relax- tigation with a controlled distribution of (10) lithologic
ation time increased as the size of the volume occupied units, an assumed state of capillary equilibrium, and
by the fluid increased. They argued that the size of a calculated the saturation level in each unit from corres-
patch (or partly drained rock joint?) may be orders of ponding capillary pressure curves. Their contrasting cal-
magnitude larger than a compliant pore. So if pore culations for shaley and clean sand, shown in Figure
fluid is arranged in patches, the apparent (Vp) stiffening 13.21, demonstrate the effect of saturation heterogeneity.
of partially saturated rock in response to a dynamic The effect of ‘patchiness’ (of saturation), compared
wave may occur even at low frequencies. to homogeneous conditions is nicely demonstrated in
Examples of frequency-dependent differences in Vp- Figure 13.22. The two parallel lines defining ‘patchy’
saturation response, for the case of carbonate samples, Vp-saturation response, are two different theoretical solu-
are shown at lower frequency (1 kHz) in the first example, tions given by Knight et al., 1998, in which just two
in Figure 13.20a. Presumably the lower frequency allows different lithologies have been combined.
the wave-induced pressure changes in the pore fluid to In Figure 13.23, the ‘extremes’ created by mixing the
dissipate, so that the pore fluid pressure is very close to ten lithological units are shown. (Properties were given
that of the high-compressibility gas in the dry pore space, in Figure 13.21a). The smooth, conventional result
as described by Knight et al., 1998. (Figure 13.23a) was obtained by a pore-scale mixture of
As a result, the pore fluid lying in thin, compliant the ten sand-to-shaley-sand units, while the ‘multi-
pores can flow freely into the dry pore space, in a squirt- stepped’ response shown in Figure 13.23b was obtained
flow type of attenuation response. It does not therefore using a patchy mixture of the ten lithological units.
Rock physics at laboratory scale 337

Figure 13.21 a) Properties of 10 lithological units, showing clay


content and permeability versus porosity. b) and
c) P-wave velocity (calculated) versus saturation for
mixtures of shaley and clean sand. Knight et al., 1998.

Figure 13.20 P-wave velocity responses to a) lower and b) higher


frequency in the case of a carbonate sample of 30%
porosity studied by Cadoret, 1993. Note ‘hardening’
at high frequency, and the different imbibition and
drainage responses with changing levels of water/air
saturation, as described by Knight et al., 1998.
Figure 13.22 P-wave velocity versus saturation in a 40% mixture
of clean sand and shaley sand. The two lithologies
13.8 Dynamic Poisson’s ratio,
were combined in a patch arrangement and as a
effective stress and pore fluid homogeneous mixture. Knight et al., 1998.

Carcione and Cavallini, 2002, described modelling


in relation to Poisson’s ratio (the dynamic value), as a The authors highlighted the fact that Poisson’s ratios
function of ‘differential pressure’ and pore fluid type. are anomalously high for cases of over-pressure, where
(The ‘differential pressure’ is the hydrostatic confine- effective stress can approach the fracturing (negative)
ment minus pore pressure, and is referred to as effective side of the usual lithostatic and pore pressure gradients
stress in soil and rock mechanics.) (as approached in Figure 12.1 of chapter 12).
338 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.24 Some experimental results for dynamic Poisson’s


ratio, as a function of effective stress (hydrostatic
confinement minus pore pressure) and fluid type.
Khazanehdari et al., 1998, as reproduced by Carcione
and Cavallini, 2002.

stiff, equi-dimensional pores does not show major vari-


ations of  with effective stress. However, closure of low
aspect ratio (compliant) microcracks and pores will tend
to increase the bulk modulus K more than the shear
modulus
, having the following effect on () :

 
1  1 
  1  
1  K 
(13.10)
2 
 3


Therefore, in dry rocks, dynamic Poisson’s ratio


increases with differential or effective stress. However, in
saturated rocks the compliant pores have become stiff-
ened in relation to high frequency waves, so () changes
less as effective stress increases. However, at low effective
Figure 13.23 P-wave velocity versus saturation for a) a pore-scale stress, or when pore pressures are very high, the effective
‘homogeneous’ mixture of ten units, b) a patchy stress sensitivity is marked, and () increases.
mixture of ten units, from sand to shaley-sand. Some experimental results that the authors quoted,
Knight et al., 1998.
from tests by Khazanehdari et al., 1998, are shown in
Figure 12.24, and indicated the increasing sensitivity at
Dynamic Poisson’s ratio, as we have seen in Part I, low effective stresses. Carcione and Cavallini were par-
can be calculated from: ticularly interested in the responses of () close to the
hydraulic fracturing limit when sealed over-pressured
1 1   Vp 2 beds were under-compacted, and where there could be
  1  , where a   
2  a  1   Vs  excess pressure due to oil-to-gas conversion, as investi-
(13.9) gated by Carcione and Gangi, 2000. As they pointed
out, at zero effective stress, Vs is (locally) zero as the
Carcione and Cavallini emphasised that it was the aspect rock mass is hydraulically fractured and load is born by
ratio of the cracks and pores and the nature of the sat- the fluid. However Vp is not zero, therefore the ratio (a)
urating fluid that determined . Rock containing mainly in equation 13.9 tends to infinity and  : 0.5.
Rock physics at laboratory scale 339

Figure 13.25 Dynamic Poisson’s ratio  versus differential stress


(hydrostatic confinement minus pore pressure). The
squares and open circles are from Winkler, 1985.
Curves marked 1, 2 and 3 were calculated results for
full oil saturation, 50% oil saturation and full brine
saturation, respectively. Carcione and Cavallini, 2002.

Carcione and Cavallini used the theory of Carcione


and Gangi, 2000, to develop theoretical trends for
Poison’s ratio increase when approaching zero effective
stress, showing in Figure 13.25, similar trends to the
available experimental results.

13.9 Dynamic moduli for estimating


static deformation moduli Figure 13.26 Non-linearity and hysteresis observed in uniaxial
cycling (with constant differential stress). The rock was
a tight gas sandstone, tested dry. Tutuncu et al., 1998.
In engineering fields that involve design for building
foundations, dams or tunnels, or indeed wellbore stability
for hydrocarbon exploration and production, it is well the 3 km of overburden, which were inevitably velocity-
documented that static moduli of deformation can be sig- based at that time. Only the reservoir itself (chalk) was
nificantly lower than the dynamic moduli predicted from core-sampled and laboratory tested.
P- and S-wave velocities. (See Chapter 6, for extensive Based on laboratory observations of the elastic non-
rock engineering based comparisons of these quantities.) linear behaviour for sandstones as illustrated in Fig-
Such differences will also tend to be greater in jointed in ure 13.26 a and b, Tutuncu et al., 1998, showed that
situ rock, as opposed to microcracked lab samples. the frequency of measurement was all important for the
Such differences in moduli are of obvious interest to geophysical estimate obtained, since Eultrasonic  Elog 
the petroleum industry. Tutuncu et al., 1998, expressed Elow freq.  Estatic. Their ultrasonic laboratory measure-
the opinion that knowledge of non-linear elastic prop- ments were conducted at 1 MHz, 180 kHz, 100 kHz
erties (that are largely responsible for the differences and 50 kHz, and their low frequency measurements at
between dynamic and static moduli) is essential for opti- 2 kHz to 1 Hz, and their static measurements at 0.05 Hz
mal drilling, effective well completions and efficient reser- to 0.001 Hz.
voir management. For example, when applying distinct To understand the frequency dependence of the vari-
element modelling to the Ekofisk reservoir subsidence ous ‘dynamic’ moduli, it was necessary to see the effect
in the mid 1980s, (Barton et al., 1986), it was appreci- of frequency on the velocities of the P- and S-waves.
ated that there were obvious uncertainties about the Firstly, for the case of dry porous sedimentary rocks, it
choice of moduli of deformation for the various layers of was generally concluded by the authors that dynamic
340 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.27 Ratios of Vp static/Vp dynamic, and Vs static/Vs


dynamic, and attenuation, each as a function of strain
magnitude. Tutuncu et al., 1998.
Figure 13.28 Young’s modulus as a function of strain amplitude at
tests gave results that were independent of frequency two different confining stress levels, and a comparison
below ultrasonic frequencies. However, when ultra- of E dynamic/E static. Tutuncu et al., 1998.
sonic frequencies were approached (0.1 MHz) the
wave length could become comparable to the grain size stiffer. By contrast, low-frequency measurements give
and scattering became an important attenuation mech- sufficient time for fluid transfer and squirt to occur
anism, which increased with frequency and when het- from microcracks, so that a relaxed, drained or less stiff
erogeneities were present. behaviour is registered. Squirt flow appeared to be the
In fluid saturated porous rocks, Vp, Vs and attenu- dominant mechanism for attenuation and velocity dis-
ation depend on frequency even when scattering is neg- persion at frequencies from 100 Hz to 10 kHz.
ligible. Since Vp increases with frequency more than Vs, When strain amplitude was increased (as in static
the resulting Vp/Vs increase may give a pronounced measurements), the good agreement between wave propa-
increase in the dynamic modulus estimate. Under the gation models and experimental data broke down, and
more rapidly oscillating loads, the fluid in the pores and large discrepancies were experienced between measured
grain boundary cracks are not allowed sufficient time to and predicted velocities and attenuations.
(micro-) flow or ‘squirt’, and the rock acts as if unre- One of the main objectives of the Tutuncu et al.,
laxed so the properties measured will be undrained and 1998, study was to develop a methodology to predict
Rock physics at laboratory scale 341

hard-to-measure static (low frequency) moduli and


attenuation from the relatively easy to measure dynamic
moduli. They therefore conducted stress cycling meas-
urements at various stresses and cycling frequencies.
The fact that Vultrasonic  Vlog  Vstatic is emphasised
by their plots of Vp ‘static’/Vp dynamic versus differen-
tial strain amplitude. A typical set of Vp, Vs, Young’s
modulus and attenuation data plotted in this format is
shown in Figure 13.27 a, b and c.
For the above general reasons, when comparing
dynamic Young’s modulus to ‘static’ Young’s modulus,
Tutuncu et al., 1998, were able to show large discrepan-
cies, since the ‘static’ measurements (0.05 to 0.001 Hz)
were at so much higher strains than low strain dynamic
measurements. Results are shown in Figure 13.28, which
resemble rock engineering data shown in Chapter 6.

13.10 Attenuation due to fluid


type, frequency, clay,
over-pressure, compliant
Figure 13.29 Compressional wave velocity and seismic quality factor
minerals, dual porosity
Qp as a function of gas/brine saturation for sandstone
at 10.3 MPa effective stress. Frisillo and Stewart, 1980.
In this section we will trace parts of the development of
attenuation as a means of improved characterization of
saturation) was matched initially, by greater attenuation
reservoir rocks. The dispersive, frequency-dependent
(a Qp of 33, reducing to approximately 9). At the far
nature of seismic Q, and the greater sensitivity of the
end of the saturation scale, when the samples become
ratio of Qs/Qp to fluid and partial saturation than Vp/Vs,
‘room dry’ and reached 100% saturation with nitrogen,
as already reviewed briefly in Chapter 10, will be illus-
the attenuation reduced sharply, and Qp reached a value
trated by interesting cases reported in both past and
of about 50. Clearly this is related to the eventual
recent literature.
absence of squirt flow with increased dryness.

13.10.1 Comparison of velocity and


attenuation in the presence 13.10.2 Attenuation when dry or gas
of gas or brine or brine saturated

Some of the subtle differences between velocity and In 1979, Toksöz et al. presented ultrasonic laboratory
attenuation (in the form of seismic Q) were shown by data on dry and water- or brine-saturated rocks, inves-
Frisillo and Stewart, 1980 tests with variable gas/brine tigating how the attenuation varied with ‘differential’
saturations. The tests on Berea sandstone (n  19.7%, pressure. Pore fluid pressure (Pf) and confining pressure
k  376 mdarcy) shown in Figure 13.29 give Vp and Qp (Pc) on their specimens was controlled independently,
on the vertical axes, and percent gas saturation on the Pc  Pf giving their quoted ‘differential’ pressures. In
horizontal axes. Zero percent gas saturation corresponds a companion paper, reviewed in Chapter 10, Johnston
to 100% brine saturation. et al., 1979, presented the assumed mechanisms of
Frisillo and Stewart’s data represented by black circles, attenuation, and formulated theoretical models that fit-
is shown supplemented by some reasonably consistent ted this laboratory data. The present data was obtained
data (also for Berea sandstone) reported by Spencer, at ultrasonic frequencies (0.1 to 1.0 MHz), using a
1979 (squares), and by Toksöz et al., 1971 data for dry Berea sandstone with 16% porosity.
or full saturation (triangles). The expected reduction Toksöz et al., 1979 used a frozen rock (limestone),
in Vp by reduced brine saturation (and increased gas showing very small attenuation and very high Qseis, as a
342 Rock quality, seismic velocity, attenuation and anisotropy

(a)

(a)

(b)

Figure 13.31 Qp and Qs values as a function of differential pressure


(Pf  0.465 Pc) for brine-saturated Berea sandstone,
with NaCl concentrations of a) 67.2 ppm and
(b) b) 161.3 ppm respectively. Toksöz et al., 1979.

Figure 13.30 Qp and Qs values as a function of confining (or dif- The anomalous high value of Qp at highest differen-
ferential) pressure, in a) dry (Pf  0) and b) methane
tial pressure in Figure 13.30a was assumed to be due to
saturated (Pf  0.465 Pc) Berea sandstone. Toksöz
some pore collapse and locking of grain boundaries.
et al., 1979.
Several important trends can be seen in the Q data.
reference standard. Their subsequent detailed studies of Firstly, Qp and Qs were both higher (less attenuation)
environmental effects such as dry, methane-saturated, in the dry or methane saturated states, than in the case
or brine-saturated, were conducted on Berea sandstone. of brine saturation. Furthermore, Qs was more often
At zero pressure, the P-wave velocities of their dry larger than Qp in the case of the dry and methane-satu-
and brine-saturated Berea sandstone were about 3.3 and rated sandstone. In contrast, there was a consistently
3.8 km/s respectively, rising rapidly to about 4.1 and wide separation of Qp and Qs of some 100 to 150%
4.2 km/s at 3000 psi (about 21 MPa). At the highest (Qp  Qs) in the case of the brine saturation.
differential pressure used there was little difference
between the dry and brine-saturated Vp in relation to
the differentiation of the dry or brine-saturated condi- 13.10.3 Effect of frequency on
tion. These and similar velocity trends for methane velocity and attenuation,
and for other brine concentrations were used in their dry or with brine
calculation of seismic Q, which are shown in Figure
13.30 and 13.31, for the case of dry, methane-satu- Winkler, 1983, provided a remarkably detailed collection
rated, and two NaCl concentrations. of velocity and attenuation data for three sandstones
Rock physics at laboratory scale 343

when tested in triaxial compression, over a range of in the case of the three sandstones could readily induce
frequencies (400 to 2000 kHz) in a dry or brine- fluid flow into and out of the contact regions.
saturated state. Effective stresses were increased in Interestingly, the reported behaviour of fused glass-
the sequence 2.5, 5, 10, 20 and 40 MPa in the case of bead samples was entirely different, as the sintered con-
the dry samples. tacts between the glass beads were very stiff and lacked
The three sandstones had the following basic phys- pressure dependence, and would not therefore generate
ical properties: local (micro) fluid flows (squirt) which would have
increased the attenuation.
Massilon sst. Berea sst. Boise sst. Figure 13.33 shows this contrasting Vp and attenua-
tion behaviour for the fused glass-bead samples, which
Porosity % 24,6 20.3 24.9
had spherical 177–210
m beads (initially) that were
Permeability (mD) 1425 107 286
Grain size (
m) 200 150–200 150–200
fused by heating to give a porosity of 26.6% and a
Quartz % 94 80 46 permeability of ‘several’ Darcies (1 Darcy  1012 m2),
i.e. it was significantly more permeable than the three
(Note: 1 Darcy  1012 m2 or 105 m/s, approximately) ‘sandstones’, yet could not generate squirt-related
attenuation.
The three ‘sandstones’ were of Mississippian age (first
two) and Pliocene (last), with geological descriptions
quartose sandstone, greywacke and arkose, respect- 13.10.4 Attenuation for
ively. Grain densities varied only from 2.68, 2.65 to distinguishing gas
2.63 gm/cm3, respectively. condensate from oil
The frequency-dependent, saturation-dependent and and water
effective stress-dependent variations of Vp and attenu-
ation (1/Q) measured by Winkler in this study, are On the basis of the differentiation of Qp values, listed
reproduced in Figure 13.32 (a to f ). below, Klimentos, 1995, suggested a way of distinguish-
The dry samples all showed negative velocity disper- ing gas and condensate from oil and water in sandstone
sion, meaning velocity decreasing with increasing fre- reservoirs, but at the same time questioned whether the
quency, while the attenuation increased as the third to method could also be used in carbonate reservoirs.
fourth power of frequency. This was taken as evidence
of scattering within the pore spaces between the grains. 1. in perfectly dry rocks, Qp is very high
(Inter-grain scattering in aluminium reportedly also 2. in fully liquid saturated rocks Qp is at an intermedi-
results in f 4 attenuation.) ate level
The brine-saturated ‘sandstones’ mostly showed 3. in partially saturated rocks Qp is low
slight, positive, velocity dispersion (at least at the lower
confining pressures) while attenuation varied with only Ranges of Qp and Qs for sandstone reservoirs were
the first or second power of frequency. This change in reportedly as follows based on well log, i.e. sonic
attenuation-frequency dependence was taken as evi- frequencies:
dence of local fluid-flow loss mechanisms. The sat-
urated rocks always showed much stronger attenuation ● Gas and gas- 5  Qp  30 15  Qs  30
(lower Qp ) than the dry samples. condensate bearing
All the data sets showed the strong influence of effec- sandstones
tive stress, which had greatest influence on attenuation ● Oil bearing 8  Qp  100 15  Qs  50
when the samples were brine-saturated and at the low- sandstones
est levels of effective stress, as we have seen earlier in ● Water bearing 9  Qp  100 15  Qs  50
this chapter. sandstones
Parallels to compliant joints affecting rock mass behav-
iour (Chapters 15 and 16), were the observations by Cross-over of P-wave and S-wave attenuation (or cross-
Winkler, 1983, that the nature of the grain contacts was over of Qp and Qs, due to the increase in attenuation of
all important. Compression and dilation of relatively P-waves by gas (lower Qp) and the absence of effects on
compliant contacts (and strong pressure dependence) the S-wave attenuation (as above), was the basis of the
344 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.32 Compression wave velocities and attenuations (1000/Qp) as a function of frequency, and whether dry (dashed line) or brine
saturated (solid line), for three ‘sandstones’. Numbers on the curves represent the applied effective pressures (MPa), Winkler,
1983. Note interpreted Qp scale added on right-hand axis for convenience.
Rock physics at laboratory scale 345

Figure 13.34 Indication of a gas zone in a well, due to Qs  Qp.


Klimentos, 1995.

A specific set of data for Qs and Qp as a function of


porosity in a gas zone in one of the wells analysed is
shown in Figure 13.34. The low Qp value for the gas-
bearing sandstone suggests the presence of partial satu-
ration with a liquid phase, and possibly higher pore
pressure as well.

13.10.5 Attenuation in the presence


of clay content

It has previously been established that compressional


wave velocities are inversely proportional to the clay con-
tent of sandstones (see previous section), with softening
of the sandstone matrix and reduced dynamic deform-
ation modulus as a result. (Han et al., 1986). In the study
of Klimentos and McCann, 1990, the first systematic
Figure 13.33 Vp and attenuation as a function of frequency,
and effective stress, for dry (dashed line), or brine-
study of intrapore clays on compressional wave attenu-
saturated (solid line) fused glass beads. Winkler, 1983. ation seems to have been made. These authors noted that
there was a general lack of correlation between porosity
and permeability for clay-bearing sandstones, but in gen-
eral low permeability was associated with high clay con-
method suggested. Typical sets of well data such as the tent, and high permeability with low clay content.
following were cited: Although the Biot, 1956a and Biot, 1956b theory
accounted well for attenuation in clay-free sandstones,
Vp km/s Qp n% it apparently failed by an order of magnitude to account
for the attenuation effect of the clay. Klimentos and
1. Gas-bearing (2.5 km depth) 4.0 5 15 McCann suggested, as others since then, that this strong
sandstone
clay-related attenuation was due to viscous interaction
2. Water-bearing (2.5 km depth) 4.2 40 15
between the clay particles and the pore fluid. Since
sandstone
the permeabilities were strongly dependent on the
346 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.35 shows the strong influence of clay con-


tent (% by volume) on Qp, and a less clear relation of
Qp with permeability. The data set excluded the few sam-
ples with zero clay content, as they could not be plotted
on the logarithmic scale.
Qp and clay content showed excellent correlation.
The best fit statistical relationship was:

Qp  179C0.843 (13.11)

where C was the percent clay content by volume. The


correlation coefficient was 0.91.

13.10.6 Attenuation due to


compliant minerals
and microcracks

A number of important trends of behaviour regarding


potential mechanisms of attenuation behaviour in the
presence of compliant minerals and joints under the
effect of confinement, have been revealed by researchers
at Imperial College, using 260 m deep research bore-
holes, located in Northern England. The near surface
sedimentary series of rocks (sandstones, siltstones and
limestones)have been extensively investigated, both in
laboratory, and with multi-frequency field surveys.
The water-saturated microcracked clean sandstones,
Figure 13.35 Top: Seismic quality Qp versus volumetric clay con-
and their in situ counterparts (bedded/jointed sand-
tent for 39 sandstones with porosities from 6 to 36%. stones) exhibited strong pressure dependence/depth
Bottom: Permeability (which is dominated by clay dependence respectively, with strong decreases in attenu-
content) versus Qp . Confining pressure was 40 MPa ation, and increases in Qp at higher pressures. (Best and
in each case. Klimentos and McCann, 1990. Sams, 1997; Best, 1997).
This effective stress dependence was attributed to
reduced squirt flow as pressure rose, due to partial closure
clay-content, the dependence of the attenuation on the of joints and bedding planes. Other sandstones and silt-
permeability arose from the over-riding influence of stones at the site had varying proportions of clay and
clay-content. One may also add here that any possible kerogenic organic matter, which seemed to reduce sen-
thin layering of clay-rich horizons could, in an in situ sitivity to pressure; Qp was quite low even at high pres-
environment, cause an increase in the shear compliance. sure, which was attributed to ‘clay squirt flow’, Best
Clay-bearing discontinuities tend to have low (‘static’) 1997. Data from Best et al., 1994 and Best, 1997, com-
shear stiffness and high (dynamic) shear compliance. bined in Figures 13.36 a and b, shows systematic reduc-
Klimentos and McCann, 1990, measured the attenu- tion of Qp (from 80 plus to about 20) as the percentage
ation of compressional waves in 42 sandstones under a of compliant minerals in sandstones and siltstones
confining pressure of 40 MPa. The frequency range was increased from a few percent to nearly 80%.
limited to 0.5–1.5 MHz, using a pulse-echo method. Best and Sams, 1997 speculated that clay squirt flow
The objective was to investigate the role of porosity, clay would be an important mechanism at both seismic and
content and permeability on the attenuation. Intrapore sonic frequencies, if larger scale geologic features were
clays were naturally found to be important in causing involved, such as inter-bedded permeable and imper-
attenuation and in modifying the permeability. meable layers.
Rock physics at laboratory scale 347

(a)
(a)

(b)

Figure 13.36 Ultrasonic data for siltstones and sandstones; Qp ver-


sus compliant mineral content, showing the effect of (b)
confinement. Best, 1997 and Best et al., 1994.
Figure 13.37 Comparative Vp, Vs and Qp, Qs responses of a sat-
urated, microcracked sample of sandstone to confin-
From a rock mechanics/rock engineering point of ing pressures varying from 5 to 60 MPa. Best, 1997.
view it would seem important to measure the deform-
ation properties imparted by these increased contents
of compliant minerals. In the Q-value world of rock confining pressures of up to 60 MPa are compared in
mass engineering quality, increased compliant mineral Figure 13.37.
content would mean reduced uniaxial strength c and The sensitivity shown alerts one to the possible com-
therefore reduced Qc (Q  c/100) and modulus. plication of sampling-induced microcracking (i.e. related
There were clearly significant differences in the elastic to the initial stress release when coring), which is ever-
deformation properties of the siltstones, sandstones and present in laboratory samples acquired from anisotropic-
limestones as intact materials. Best, 1974, mentioned a stress environments, if tested at artificially low confining
maximum range of instantaneous sample shortenings of pressures.
0.74% (siltstone), 0.18% (sandstone) and 0.05% (lime- At 5 MPa confining pressure (close to the estimated
stone) as a result of applying the 60 MPa confining pres- in situ confining pressure at the borehole site), Qp was
sure. Moduli of 8.1, 33.3, and 120 GPa are implied. 24  2, increasing to 83  29 at 60 MPa. The grain
P-wave velocities for the laboratory samples were about contact microcracks were apparently closing beyond
3.4 to 3.7, 3.6 to 4.6 and 6.0 to 6.3 km/s respectively. about 40 MPa, with a consequent reduction in squirt
In the case of a clean, saturated sandstone pervaded flow related attenuation.
by microcracks, Best and Sams, 1997 and Best, 1997, The most interesting set of data from the point of
showed that Qp was a more sensitive indicator of the view of rock mass quality (as opposed to rock matrix qual-
effect of confining pressure than Vp. The compara- ity), and in relation to attenuation behaviour is shown
tive ultrasonic responses of velocity and attenuation to in Figure 13.38. This strictly field data (belonging to
348 Rock quality, seismic velocity, attenuation and anisotropy

for velocity-depth gradients to 250 m that are in excess


of 5 s1, that Q c  0.1.
A typical finely-layered sequence of limestones, sand-
stones, siltstones and mudstones, with bedding and
perhaps a set of bed-limited joints, could be expected to
have a ‘poor-quality’ rock quality Q-value, lowered by
bedding planes (counted as one set, as is customary, and
possible cross-bedding, giving a moderate RQD. Softer
inter-beds would effectively increase Ja (See Appendix A
for a description of these terms.)
We may estimate Q  20/4  1.5/4  0.66/1, or
about 1, certainly to the nearest order of magnitude.
Together with a rough estimate of uniaxial strength of
closer to 10 to 40 MPa than either 1 or 100 MPa, the
resulting Q c value would be very roughly 0.1–0.4,
making velocity of about 3.5 km/s at about 250 m
depth easily understood, if the matrix porosity was
about 10% (refer to Figure 11.73).
A medium low rock quality Q-value, typical for sedi-
mentary inter-bedded strata, and a correspondingly
Figure 13.38 Comparison of core (Whitchester sandstones) and
lower rock quality Q c value, would give significantly
downhole sonic log measurements, at equivalent
reduced rock mass static modulus of deformation Emass
pressures and depths. Best and Sams, 1997.
(or M), Barton, 1995, from the relation:
M  10  Q 1/3
c
(13.12)
Chapter 14) is presented here for the sake of continuity.
Presumably an ‘altered zone’ (stress-fracturing and mud- Values of M would be in the range 4 to 7 GPa, for the
filtrate invasion: Chapter 12) that tends to affect sonic case of Q c  0.1–0.4. This ‘static’ modulus actually has
logging data at reservoir depths, would be only a very the appearance of correlating, in broad terms, with the
minor factor at this shallow 260 m deep well. lowest values of seismic Q p obtained at lowest frequency.
The figure shows the stress dependence of the micro- In Chapter 10 we saw frequency-dependent seismic Q
cracked laboratory specimen (mild) and the stronger values as low as 9 to 12 for this in situ site (Figure 10.35),
response to pressure of the bedded/jointed in situ struc- from sonic log based (8 to 24 kHz) measurements.
tures. Best and Sams, 1997 did not comment on the
relative bedding/jointing frequency at the different
depths in the borehole. However, the pressure/depth- 13.10.7 Attenuation with dual
velocity gradient shown in Figure 13.38 can be estimated porosity samples of
to be approximately: limestones

3.6  2.7 (km/s) 0.9 Assefa et al., 1999, conducted ultrasonic (0.7 to
  10 s1 0.85 MHz) compressional and shear wave attenuation
3.0  0.8 (MPa) 0.09 measurements on forty, 5 centimetre-sized samples of
water- or oil-saturated oolitic limestones, at 50 MPa effec-
In the above, the pressure increment 2.2 MPa has been tive hydrostatic confining pressures ( '  55 – 5 MPa
converted to an approximate 90 m depth (0.09 km) pore pressure  50 MPa). They found that attenuation
to give the conventional Vp-depth gradient in s1 units. reached a maximum value in the samples which had the
This steep gradient suggests quite a low rock quality most fully developed ‘dual-porosity’. This dual porosity
Q-value, representing the soft porosity in this bedded consisted of inter-particle macro-pores (dimensions up
sedimentary strata, which responds strongly to pressure to 0.3 mm) and micro-pores (with dimensions of 5 to 10
increase at shallow depth. In Chapter 11, Figure 11.72, microns). They interpreted this increased attenuation
an empirical Vp-Q c-depth trend was shown, that suggests, (low Q seis) as evidence of a squirt-flow mechanism, as
Rock physics at laboratory scale 349

Figure 13.40 Compressional-wave seismic quality factor (Q p) versus


shear-wave quality factor (Q s), showing a dominance
of Q p over Q s when there was less attenuation. Assefa
et al., 1999.

was about 5 in water/oil saturated sandstone. The


combination of Q p/Q s and Vp/Vs with other well logs
enabled differentiation of gas-bearing from oil-bearing
reservoirs, as we have seen in other research.
Assefa et al., 1999, found that their ‘dual porosity’ (or
bimodal porosity) limestone specimens showed higher
attenuation (lower seismic Q) when permeability and
Figure 13.39 Attenuations (1/Q p and 1/Q s) for water saturated (total) porosity were also larger. Their results are shown
limestones, as a function of (helium) porosity (range in Figure 13.39. Both distributions of pore size were
3 to 17 %) and a roughly three-orders-of-magnitude important, and the attenuation was shown to be the sum
permeability scale. Assefa et al., 1999. of Biot-type fluid flow and squirt flow to/from the larger,
moderately interconnected inter-granular pores, which
found in shaley sandstones. They also suggested that sometimes contributed about 90% of the total porosity.
conventional dual porosity (i.e. joints and pores) present Assefa et al., 1999 also presented results for the ratio of
in the case of, for instance, in situ limestones, could sim- Q p and Q s, sorted by mineralogical differences in their
ilarly cause seismic attenuation due to squirt flow. samples. Figure 13.40 shows the generally larger magni-
Modelling these dual porosity aspects will be addressed tude of Q p consistent with the general effect of saturation,
in Chapter 15. as referred to above. They also compared Q seis values (Q p
Klimentos and McCann, 1990, and others had pre- and Q s) for oil-saturated and water-saturated, showing in
viously shown how attenuation in sandstones was Figure 13.41 how the water saturated specimens generally
dependent on the pore-filling minerals, particularly the showed less attenuation (higher Q p in particular).
clay content. They had shown how the attenuation was The authors posed the question of whether the
significantly higher in clay-rich sandstones, than in clean ultrasonic data (0.7–0.85 MHz) for these small ‘intact’
clay-free sandstones of identical porosity. Klimentos, dual-porosity limestone samples, were of any value to
1995, had later measured compressional- and shear-wave geophysicists trying to interpret propagation through
attenuation from sonic wave forms in three gas and dual-porosity (porous and jointed) limestones in the
oil reservoir wells, and was able to show that Q p/Q s field, at frequencies in the 50 Hz to 30 kHz ranges, as
was about 1/3 in gas bearing sandstone, while Q p/Q s used in seismic and sonic log surveys.
350 Rock quality, seismic velocity, attenuation and anisotropy

potential hazard due to the risk of so-called shallow water


flows (SWF). Sands in the Gulf of Mexico can present
problems for these reasons, at water depths of between
400 and 2100 m, and depths below the mudline of as
much as 1200 m. As the name implies, SWF can also be
a hazard in shallow water drilling, where effective stresses
and compaction of sediments can be minimal, and pro-
gressive instability during drilling at a new well can
potentially engulf neighbouring wells.
Prasad, 2002, used rock-physics principles involving
velocities and attenuation, to study this problem. Older
data on sands tested at very low effective stresses (in
Prasad, 1988), were added to by new data, by perform-
ing 1 MHz pulse generator testing of jacketed, lightly
confined, clean beach sands of grain size 250–550
m.
Due to very low values of shear wave velocity at low
effective stresses, for example 400 m/s at 1 MPa, while
Figure 13.41 Comparison of oil-saturated and water-saturated equivalent compression wave velocities were closer to
seismic quality factors, for bi-modal porosity lime- 1,800–1,900 m/s, there was an exponential increase in
stones, at 50 MPa effective hydrostatic stress. Assefa the ratio of Vp/Vs to values beyond 5 and 10, and even
et al., 1999. beyond 100 at negligible effective stress. This trend,
which became very noticeable below 2 MPa, is shown
Concerning in situ jointed, dual-porosity chalk (with
in Figure 13.42a, with a log scale for effective stress
higher porosity than the limestones), Newman and
shown in Figure 13.42b.
Worthington, 1982, measured Q p and Q s values as low
Prasad showed that there was a dramatic change in
as 4.0 and 5.2, and 3.5 and 5.9, in two near-surface fis-
the S-wave signals when testing at extremely low effect-
sured (jointed) chalks, using seismic frequencies. Assefa
ive stress levels, with high attenuation of the shear waves
et al., suggest that when interpreting the propagation of
at the lowest pressures, indicating the sand was close to
low frequency seismic waves, the potential for squirt
a state of suspension, with low shear strength. The
flow attenuation in a large scale dual porosity system
amplitude of the S-waves decreased dramatically below
such as jointed limestones will clearly be present.
1 MPa. In contrast, P-wave attenuation reduced mar-
Possibly this is why a certain correlation is being
ginally at very low effective stress. Due to the decreasing
noticed between seismic Q and the ‘static’ modulus of
S-wave velocity, Poisson’s ratios increased rapidly to just
deformation, expressed in GPa and readily estimated
below 0.5 at negligible effective stress levels.
from rock quality Q (equation 13.12). The components
The diagnostic use of the velocities and seismic Q
of the rock quality Q-value reflect many potential
values for registering over-pressure, and the presence of
attenuation-causing factors, e.g. RQD/Jn for scattering
gas can conveniently be ‘tabulated’ as follows, based on
due to relative block size, Jr/Ja concerning frictional and
the author’s summary text:
conductive properties of the joints expected to be rele-
vant for squirt flow (also including mechanisms in
clay), Jw as a direct link to permeability, and SRF related Detecting over-pressure
to increased attenuation where stress is low, and reduced Reduction in Vp may be ambiguous, as both over-pressure and gas
attenuation where stress is high. reduce Vp
Vp/Vs increases with over-pressure
Vp/Vs and Poisson’s ratio both increase exponentially, when
13.10.8 Attenuation in the presence
sediment approaches a state of suspension
of over-pressure
Detecting gas
Reduction in Vp may be ambiguous, but Vs will be unaffected by
Establishing wells in deep sea environments, where there
presence of gas
may be over-pressured zones due to rapid sedimentation
Vp/Vs decreases with gas saturation
of alternating sands and shaly sediments, presents a
Rock physics at laboratory scale 351

Figure 13.43 Distinguishing between sand and sandstone is very


clear using the ratio of Q p/Q s in combination with
(Vp/Vs)2. The effect of increasing pore pressure, and
reducing effective stress is suitably accentuated.
Prasad, 2002.

They pointed out that rocks saturated with fluids of


high compressibility and low thermal expansion coeffi-
cient were generally under-pressured, while rocks sat-
urated with fluids of low compressibility and high
thermal expansion coefficients were generally over-
pressured. The latter could therefore be seismically ‘visi-
ble’. Of course at high differential pressures, the veloci-
ties (and quality factors) became almost constant. The
authors’ model was able to predict pore pressure from
seismic properties, if reliable wave velocities and quality
Figure 13.42 Ratio of Vp/Vs showing exponential increase at low factors could be obtained.
effective stresses (Note: differential pressure Pd is effect-
ive stress, defined as Poverburden – Ppore). Prasad, 2002.
An empirical prediction from work by Hamilton, 13.11 Attenuation in the presence
1971 is glimpsed beneath the recent data in the upper of anisotropy
diagram.
A further combination of interesting rock physics test
Carcione and Gangi, 2000, added another important data and sophisticated modelling abilities were described
aspect to the understanding of over-pressure effects on by Carcione, 2000, concerning petroleum source rock
seismic attenuation, by considering and modelling, containing kerogen and different amounts of water. The
the relative effects of pore-space compressibility, and the objectives of the study were to obtain a model for source
compressibility and thermal expansion coefficient for the rocks that would be capable of relating seismic anisotropy
fluid mixture filling the pore space. Their model con- (of velocity and attenuation) to kerogen content, pore
sisted of reservoir sand that was buried at a constant pressure and water saturation. The author succeeded in
sedimentation rate, under a geothermal gradient which demonstrating that anisotropic velocities and attenuation
was constant both in time and depth. Their model could be used as strong indicators of kerogen content and
(Figure 13.44) showed realistic reductions of velocity maturation, which depends on pressure change.
and quality factor with reducing differential (or effect- Some test data, from the North Sea Kimmeridge Shale,
ive) pressure, especially when less than 15 to 20 MPa, as is shown in Figures 13.45a and b. The author’s model,
commonly observed in numerous studies reviewed in shown as solid curves, was based on a viscoelastic trans-
this book. versely isotropic medium composed of illite and smectite
352 Rock quality, seismic velocity, attenuation and anisotropy

(a) (b)

(c) (d)

Figure 13.44 Modelling of low frequency (25 Hz): a) and b), and ultrasonic (1 MHz): c) and d) P-wave seismic Q versus differential pressure,
or versus excess pressure, for a water-saturated model sandstone reservoir. Experimental squares for bedding-parallel Q of Berea
sandstone, from Prasad and Manghnani, 1997. (Dotted lines correspond to 0 to 2 km, where the rock is normally pressured, and
the continuous lines to the range 2 to 8 km where the rock is over-pressured.) Carcione and Gangi, 2000.

and organic matter. The data for Vp and Vs versus kerogen compared to the calculated lithostatic minus hydrostatic
content, was presented from measurements in bedding pressures of 82 MPa and 34 MPa. A kerogen content of
normal (0°) and bedding-parallel (90°) directions. The 35% was assumed here. The sonic log result for Vp
marine Kimmeridge Shale was from the Draupne showed a typical reduction from about 4 km/s to a fairly
Formation, located at between 3480 and 3580 m depth, constant value of only 2.6 km/s for the 100 m thick sec-
in the central Viking Graben of the North Sea. tion of this valuable over-pressured source rock.
The modelled seismic Q p and Q s trends, shown in dia-
grams d) and e), demonstrated an attenuation anisotropy
that was higher than the velocity-based stiffness 13.11.1 Attenuation for fluid front
anisotropy. The largest anisotropies were at 18% kerogen monitoring
content for the case of attenuation, and 30% kerogen con-
tent for the case of stiffness or velocity. In Figure 13.45, In 4-D, time-lapse seismic monitoring of reservoir
diagrams e) and f ), Carcione demonstrated the modelling processes, several 3-D seismic surveys of the same reser-
of Vp and Vs for the case of a fixed kerogen content of voir locations made at different times are compared.
35%, and several water saturation levels (Sm  0 to 0.7), Differences in reflection amplitude or impedance indi-
again for the case of bedding normal (0°) and bedding par- cate changes in the reservoir. There is an increasing move
allel (90°) directions. A key variable in these plots was the to have permanent down-hole sources and receivers for
strong effect (in the case of the bedding-normal (0°) direc- obtaining greater detail of the movement of fluids dur-
tion of measurement), of an excess pore pressure of up to ing production and injection (Ziolkowski, 1999), and
50 MPa, giving as expected, successively lower velocities, for enabling the periodic performance of high fre-
with greatest reductions for the highest saturation levels. quency cross-hole imaging (e.g. see Chapter 14).
The different maturation stages of this source rock were Wulff and Mjaaland, 2002 studied time-dependent
modelled by evaluating the kerogen to oil conversion fluid-front 4-D seismic effects, using a large scale
and the excess pore pressure, with fracturing estimated if laboratory test, in which a block of lower Triassic sand-
a change of pore pressure of as much as 48 MPa occurred, stone of 17% porosity, was successively submerged in a
Rock physics at laboratory scale 353

(a) (b)

(c) (d)

(e) (f)

Figure 13.45 Experimental velocity data for Kimmeridge Shale source-rock samples from three and a half kilometres beneath the North Sea,
with comparison to viscoelastic, transversely isotropic model results. Anisotropic velocity, attenuation, and velocity depend-
ence on excess pore pressure was demonstrated, when measured or modelled in bedding-normal (0°) and bedding parallel
(90°) directions. Carcione, 2000.

water tank in four stages, with immediate and long 1/Q s of 0.3. The authors therefore preferred to use the
term (300 days) seismic monitoring of the effect of the attenuation calculated from (). This gave 1/Q p of 0.065
intermittent ‘water-flood’ and capillary effects. They and 1/Q s of 0.07. Attenuation had increased very rapidly,
used P- and S-wave transducers, with centre frequency such that maximum attenuation was reached prior to full
of 500 kHz. Transmitted and reflected waves were used saturation, followed by a quite fast decline in the three fol-
for the monitoring, with six transducers glued to the lowing months. This was assumed to be due to the
top and (submerged) bottom of the block. decline in local fluid flow (squirt) effects which were
The water-flooding caused the velocity, amplitude dominant prior to full saturation, prior to full flooding.
and frequency of the transmitted waves to diminish The authors assumed that the velocity reductions were
significantly, with reversal upon final drying of the block. due to water adsorption effects causing a reduced mod-
The authors evaluated both the attenuation coefficient ulus (strictly a laboratory, as opposed to an in situ reser-
(), where 1/Q  V/f, and the attenuation 1/Q, rela- voir effect, unless gas replaced by water-flooding could
tive to the signals obtained for the dry rock. Seismic be considered?). They concluded that improved inter-
Q was also estimated by the spectral ratio method. pretation of reservoir processes required not only P- and
At maximum saturation of the block, the spectral ratio S-wave measurement, but also amplitude and attenua-
method indicated a very high maximum 1/Q p of 0.5, and tion measurement.
354 Rock quality, seismic velocity, attenuation and anisotropy

13.12 Anisotropic velocity and just as it is for the case of rock joints, as reviewed in
attenuation in shales Chapter 3. Figure 13.47 shows the complete distribution
of velocities, with the vertical/horizontal axes showing
Anisotropy in shale has not been frequently studied due respectively the perpendicular to bedding and parallel to
to difficulties with sample disturbance when handling bedding magnitudes (units km/s).
fissile materials. The set of data from Johnston and Domnesteanu et al., 2002, measured the anisotropic
Christensen, 1995, (Figure 13.46), shows the effect of velocity and attenuation of fully saturated shales under
Vp measurement direction, and the effect of confining overpressured conditions, Figure 13.48, apparently for
pressure, and is a useful summary of the effects of pre-
ferred clay mineral orientation.
Data sets reviewed by Johnston and Christensen, 1995
showed Vp anisotropy of 20 to 30%, and Vs anisotropy
of 32 to 35%. The maximum velocity was always parallel
to bedding and the minimum perpendicular to bedding,

Figure 13.47 Velocity anisotropy caused by the preferred orienta-


Figure 13.46 Effects of confining pressure and direction of veloc- tion of clay particles in the fabric of the shale: sam-
ity measurement for shale. Anisotropy is caused by ple New 7. Note shear-wave splitting when making
the preferred orientation of clay particles in the fab- an increasingly acute incident angle to the direction
ric of the shale. (Vsh is the velocity of the shear wave of bedding. (Vsh is the velocity of the shear wave
vibrating parallel to bedding, and Vsv is the velocity vibrating parallel to bedding, and Vsv is the velocity
of the shear wave vibrating perpendicular to bed- of the shear wave vibrating perpendicular to bed-
ding). Johnston and Christensen, 1995. ding). Johnston and Christensen, 1995.
Rock physics at laboratory scale 355

Figure 13.48 a) Propagation and vibration directions with respect to foliation, and relevant to shale sample number. b) Seismic qualities for
P- and S-waves through over-pressured shales, as a function of differential pressure, and as a function of propagation direction
relative to the foliation. c) One set of examples of P-wave and S-wave attenuation as a function of over-pressure, where
PP/Pc  0.46 corresponds to over-pressure. Domnesteanu et al., 2002.
356 Rock quality, seismic velocity, attenuation and anisotropy

the first time in the laboratory. They used an ultrasonic waves propagating parallel to the layering (Vp11). Q p
reflection technique. The shale cores were from the reduced with overpressure, regardless of propagation
North Sea. Confining and pore pressures were applied direction, while Q s was highly dependent on propaga-
that were relevant to in situ conditions, giving differen- tion direction, and depended less on pore pressure than
tial pressures from 5 to 60 MPa. (In view of the low per- Q p. Perhaps predictably, the results suggested a strong
meability of the shale, the use of the rock mechanics link between the rock framework, the pore geometry and
adjective for 1  3  differential stress, will not be connectivity, and therefore the response of pore fluid to
questioned, as the actual effective stress is uncertain, due the propagation of seismic waves, in specific directions.
to test rate dependency).
The seismic signature of the shale was explored by
taking the samples through cycles of over-pressured and 13.12.1 Attenuation anisotropy
normally pressured states, whilst increasing the overall expressions E, G and D
confining pressure. The authors found that each incre-
mental increase in pore pressure caused the shale to At a seismic Q workshop in Madrid, Best et al., 2005,
expand slightly, counteracting the opposite effect of described an ultrasonic pulse-echo technique for investi-
increasing the confining pressure. It was found that the gating both velocity and seismic Q anisotropy of P- and
shales behaved elastically at confining pressures higher S-waves in finely inter-bedded reservoir-type rocks. They
than 35 MPa. As expected, the compressional and shear had modified the earlier version of the equipment, so as
wave velocities and seismic quality factors increased to be able to study shear wave anisotropy, using a 360°
with increasing differential pressure (i.e. with reduced rotating S-wave transducer, for observing the shift of
over-pressure. arrival time. This will be described here because of obvi-
The plane of circular symmetry was parallel to the ous relevance to attenuation in finely bedded reservoir
foliation (see Figure 13.48a). The degree of sample rocks, which will be partly addressed in the next section.
anisotropy was found to be related to the depth of origin Their studies were performed at effective stress levels
of the shale. Both Vs and Vp decreased with increasing of 5 and 40 or 50 MPa, on vertically or horizontally
over-pressure. Nevertheless, the ratio Vp/Vs decreased aligned samples of Carboniferous sandstone, siltstone
with increasing differential pressure. As expected there or limestone obtained from the Imperial College exper-
was a general increase in seismic Q p and Q s with differ- imental borehole site in northern England.
ential pressure, but with lowest values showing least Following Thomsen (1986) velocity anisotropy
increase with stress. Under a fixed differential pressure, expressions for ,  and  in weakly transversely isotropic
an increasing pore pressure was found to reduce Q p by media, and because of the excellent stability of their
about 16%, or up to 8 units. An anisotropic pattern of pulse-echo data, they gave equivalent expressions for
wave attenuation is evident from Figure 13.48b. attenuation (Q1), and derived relevant results for
The authors found that the relative proportions of these new parameters:
Biot ‘fluid-past-frame’ attenuation, and local squirt
flow attenuation were different in the plane parallel to 1 1
Q p (H )  Q p ( V )
foliation, and in the plane perpendicular to the layer- (Q 1 )  1
(13.13)
Q p (V )
ing. Squirt flow, localized between compliant and non-
compliant pores, was considered a predominant loss in
the bulk modulus, and a small loss in the shear modu- 1 1
Q sh (H)  Q sh ( V )
lus. It was naturally considered to be more active in the (Q 1 )  1
(13.14)
plane of the foliation, than perpendicular to it. Biot Q sh ( V )
flow was considered a predominant loss in the shear
modulus, and a very small loss in the bulk modulus. 1 1
Q sv (H)  Q sv (H)
Since arising from fluid movement in the open pores, it ((Q 1 )  1
(13.15)
was considered to be related to the macro-permeability. Q sv (H)
The authors found that at differential pressures
20 MPa, compressional waves propagating perpendi- where V denoted vertical, H horizontal, and the
cular to the layering (Vp33) were attenuated by squirt flow, sub-scripts p, sh and sv denoted P-waves, S-waves with
and hence more attenuated than the compressional horizontal polarization, and S-waves with vertical
Rock physics at laboratory scale 357

shear wave splitting for revealing attenuating and fluid-


conducting joint structure, in hydrocarbon reservoirs.

13.13 Permeability and velocity


anisotropy due to fabric,
joints and fractures

A very interesting laboratory study of velocity and per-


meability anisotropy, for the case of tight gas sandstones
Figure 13.49 Orientation of laboratory (sub-core) samples, with
containing sub-vertical, conjugate-type jointing, was
bedding features (dashed lines) either perpendicular
or parallel to the shear wave transmission. Shear
reported by Dürrast et al., 2002. The individual joints
wave polarization is shown by the Sv and Sh compo- were either ‘open’, open-and-mineralized, or mineralized.
nents. Best et al., 2005. The sandstones were of very low matrix permeability
(30
d), and had a sedimentary layering consisting of
fine clay layers. Production from a reservoir in such
Table 13.2 Q 1 anisotropy parameters for thinly bedded sandstone tight sandstones naturally depends on the jointing (i.e.
samples. Best et al., 2005.
the natural fractures).
Pressure (Q1) (Q1) ((Q1) The authors were able to study the three-dimensional
Lithology (MPa) % % % P-wave velocity of spherical samples machined from the
Sandstone 5 11.8 0.0 27.3
core, in at least 100 directions, to obtain Vp symmetry
(/3.3) (/1.2) (/1.0) without prior assumptions.
Sandstone 40 42.5 75.0 7.7 Some selected results of the P-wave anisotropy meas-
(/5.1) (/4.6) (/1.9) urements, plotted stereographically on lower-hemisphere
Schmidt net projections (perpendicular to the core
axis), are shown in Figure 13.50a. Their results for six
polarization (see Figure 13.49). The authors introduced spherical samples, under the first condition of zero con-
the parameter ␨, to describe the anisotropy between the finement, showed a range of maximum Vp (saturated) of
fast and slow S-wave polarizations in the H sample. 4.6–5.2 km/s and minimum Vp (dry) of 3.7–4.0 km/s,
(The subjects of shear wave splitting and polarization are with the difference plots (saturated minus dry) showing
treated in detail in Chapter 15). a maximum range of 0.51.1 km/s for the six samples,
A brief sampling of Best et al., 2005 results for sand- under zero confinement. The sedimentary layering and
stone, using these new expressions, is given in Table 13.2. cross-bedding tended to dominate the Vp distribution,
The samples exhibited visible sub-millimetre, fine hori- with jointing (open or mineralized) having less effect
zontal layering of clay/organic matter. At an equivalent on Vp, but tending to change the symmetry of the Vp
500 m depth (for the case of low density, porous rock), distribution, giving a more monoclinic symmetry.
the effective stress of 5 MPa indicates that there is sig- Permeability and P-wave velocity measurements were
nificant Q 1 1
p anisotropy (), none for Q sh (␥), and performed on spherical samples of the same sandstones,
1 some containing both the sedimentary fabric and the
strong anisotropy for Qsv (␨), i.e. consistent with trans-
verse anisotropy. steeply dipping jointing. Vp under four levels of con-
At 40 MPa, roughly equivalent to effective stresses finement, up to 100 MPa are shown in Figure 13.50b.
at 4,000 m depth in porous sedimentary rock masses, These tests demonstrated several important trends of
there was a dramatic increase in anisotropy for Q p1 and behaviour. Naturally, in the absence of jointing, the
1
Q sh (the latter perhaps surprising), while the Q 1 sv higher permeability and velocity values were recorded
anisotropy reduced, as would be expected at higher in the plane of the nearly horizontal layering, perpendi-
stress. A second sample showed significantly different cular to the core axis. Figure 13.51a shows the perme-
1 ability in three orthogonal directions, as a function of
results, except in the case of Q sv behaviour (reduction
with increased stress). confining pressure, with results dominated by the
The anisotropy of seismic Q will be addressed in much successively reducing porosities of the five selected
more detail in Chapter 15, due to the importance of samples. Note the low values of the vertical (Z-axis)
358 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.50 a) Three-dimensional P-wave velocity measurements on spherical samples in more than 100 directions, in the dry and sat-
urated states. Two of the six results (k3 and k5) for the tight gas sandstones are shown. These illustrated cases were under zero
confinement. b) Confined tests with sedimentary layering (ss) and sometimes jointing (r1 and r2), provided an anisotropic, con-
fining pressure-dependent mix of effects on velocities. (Note: sample porosities are listed in Figure 13.51). Dürrast et al., 2002.

permeabilities in each case, due to the thin dark-clay The authors commented on the significance of the sur-
layers obstructing vertical flow. face roughness of the joints providing a significant fluid
However where discrete jointing occurred, with sub- path, even at higher confining pressures. From 50 to
vertical orientation, as in the case of sample k5, the high- 80 MPa confining pressure, corresponding to depths of
est permeabilities were recorded parallel to the core axis, several kilometres, the permeabilities parallel to the frac-
i.e. vertically. This clear trend is shown in Figure 13.51b. tures were up to nine times higher than those parallel to
With confining pressure below 50 MPa, there was a the sedimentary layering. In view of the possibility of
mono-clinic symmetry of the P-wave velocity distribution conjugate shearing of such joint sets, one may speculate
caused by the combination of sedimentary layering, joint- that in a reservoir environment, with anisotropic stress,
ing, and cross-bedding, while at higher confinement, the there would be a possibility of relative maintenance of
sedimentary layering was virtually the only remaining fea- joint permeability despite high effective stresses. These
ture, resulting in a more transversely isotropic symmetry. possibilities, and the influence of joint-surface related
Rock physics at laboratory scale 359

paper, by one of the above co-authors, from the Institut


Français du Pétrole, see Rasolofosaon and Zinszner,
2002. The new equipment, based on that developed by
Arts et al., 1996, now allows independent application
of pore pressure and confining pressure, while measuring
P-wave velocities in multiple directions as we have seen.
These two authors described interesting comparisons of
the anisotropic 3D permeability and elasticity tensors of
various reservoir rocks. The permeability tensors were
obtained by a tracer injection and X-ray technique.
They found that in some cases, the elastic property
anisotropy, and the hydraulic anisotropy were closely
related in terms of symmetry directions – this occurred
when the two mechanisms shared the same cause, such
as layering or jointing. Good agreement was seen between
the two types of anisotropy for a North Sea sandstone
of 16% porosity (Figure 13.52), while in the case of
a dolomite of 23% porosity, there were marked differ-
ences due to the influence of small-scale disconnected
fissures. Cases where there was no correlation, highlighted
the challenges faced in estimating permeability and
(a) monitoring fluid flow from seismic measurements in
the field. In an earlier study with the same equipment,
Rasolofosaon et al., 2000 had shown a comparison of
crack and fabric analysis with inversion of the multi-
directional ultrasonic data.

13.13.1 Seismic monitoring of


fracture development
and permeability

An important experimental testing facility at Imperial


College, described by King et al., 1995, Shakeel, 1995,
(b)
and King, 2002, allows the application of extremely high
(hundreds of MPa) polyaxial stress states to small (40 to
Figure 13.51 a) Orthogonal permeabilities of five spherical samples, 50 mm) cubical specimens of rock. Figure 13.53 illus-
having successively reducing porosities. Vertical perme- trates the loading frame and principal loading and vel-
ability (Z) is least due to dominance of sub-horizontal ocity measurement directions, with definition of the nine
dark-clay layers. b) Orthogonal permeabilities of
components of P- and S-wave velocity, and the vibra-
jointed sample k5, which had a porosity of 10.2%.
tion directions of the S-waves. Pietzo-electric transducers
Note effect of sub-vertical jointing on the higher verti-
cal (Z) permeability in this case. Dürrast et al., 2002.
were used to produce and detect the pulses of compres-
sional waves (450–800 kHz), and either of two shear
waves (350–750 kHz), polarized at right-angles, and
properties, like roughness (JRC), and wall-strength propagating in one of the principal stress directions.
(JCS), in assisting permeability maintenance (unless each By holding the minimum principal stress very low
are too low), are mentioned in Chapter 14 and addressed (e.g. 2 or 3 MPa), and increasing 1 and 2 in unison, to
in detail in Chapter 16. high levels, it was possible to create a set of closely spaced
The large, spherical-sample test equipment used in extension/tension fractures perpendicular to the min-
the above studies, was also described in a companion imum stress direction. Shakeel and King, 1998 and
360 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.52 Comparison of elastic anisotropy, measured on a spherical North Sea sandstone sample, and hydraulic anisotropy, measured
by a tracer-injection X-ray technique in the same rock sample. ISO: isotropic, TI: transversely isotropic, MON: monoclinic
symmetry, ORT: orthorhombic symmetry. Rasolofosaon and Zinszner, 2002.

King, 2002 described dynamic hydro-mechanical (ultra- loading. At approximately 100 MPa, the set of ‘biaxial’
sonic and flow) measurements both before fracturing, extension fractures started to develop (point F), and the
during fracturing, and in load-unload cycles after perpendicular velocities Vp2 and Vs2 indicated rap-
fracturing. idly declining velocities due to the presence of the new
Figure 13.54 reproduces some of the very interesting fractures.
experimental results of principal P-wave and S-wave Figure 13.55 indicates what happened when the newly
velocities that were recorded during the initially almost fractured cube was reloaded with equal hydrostatic stress.
biaxial loading of a 6% porosity, and almost isotropic There was a more rapid increase in Vp2 and Vs2 as the
sandstone up to fracturing. Diagrams a) and b) show Vp1 fracture set was (nearly) closed, while parallel with the
and Vs1 increasing steadily parallel with the high ‘biaxial fractures, the velocities Vp1 and Vs1 behaved almost as
Rock physics at laboratory scale 361

(a)

(b)

Figure 13.53 a) Section through the 3D static and dynamic loading facility at Imperial College. Permeability can also be measured in the
presence of a high pore pressure. b) Principal loading and velocity measurement directions, with definition of the nine com-
ponents of P- and S-wave velocity, and the vibration directions of the S-waves. King et al., 1995, King 2002.
362 Rock quality, seismic velocity, attenuation and anisotropy

(a) (a)

(b)
(b)
Figure 13.55 Ultrasonic P- and S-wave components from tests
Figure 13.54 Ultrasonic P- and S-wave components from tests performed across a 6% porosity, isotropic cube of
performed across a 6% porosity, isotropic cube of Crossland Hill sandstone, when re-loaded in a polyax-
Crossland Hill sandstone, when loaded in a polyaxial ial facility in parallel and perpendicular-to-fracturing
facility. Pre- and post-fracturing results are shown, and directions, upon subsequent hydrostatic loading to
compared with excellent modelling results, from 90 MPa. Note the almost unchanged velocities meas-
Shakeel, 1995. Measurable permeability (1 md) ured parallel to the fracture set that was developed.
parallel to the pending fractures, did not develop until King, 2002.
velocities reduced at fracturing point F. King, 2002.

less than 5, somewhat resembling a reduction in defor-


before. The very closely matching analytical modelling mation modulus, when expressed in units of GPa.
results by Shakeel, 1995, using a method developed for Figure 13.56b shows the hydrostatic loading test phase
a transversely isotropic cracked medium by Nishizawa, for the same set of fractures, with a low stress seismic Q
1982 are virtually identical with the experimental result, increasing from 10 to about 29 with hydrostatic (there-
in terms of velocities. fore normal) loading to 70 MPa. The ‘before-fracturing’
King et al., 1997 reported several interesting results result is also shown for comparison, indicating seismic Q
concerning the attenuation of various of the nine velocity increasing from 20 to 40, as a result of loading to 70 MPa.
components (calculated by the spectral ratio technique), These again resemble increase in moduli, expressed in
in an equivalent series of three test cycles on a cube of units of GPa. In Figure 13.57, the attenuation for P- and
Penrith sandstone of 13% porosity. Figure 13.56a shows S-waves in the same perpendicular ‘3’ direction, are plot-
selected results for the ‘3’ direction, perpendicular to ted against permeability measured parallel to the fracture.
the pending fracture set, which presumably started to King, 2002 described the less demanding development
initiate at point F, when the major stress was increased of a set of parallel (cleavage) fractures in slate, simply by
beyond 100 MPa to initiate fracturing. Note the reduc- axially loading a cylindrical sample in a conventional
tion of seismic Q from about 20 before fracturing, to triaxial cell. These easily formed, smooth, planar cleavage
Rock physics at laboratory scale 363

(a)

Figure 13.57 Attenuation for P- and S-waves in the ‘3’ perpendi-


cular to fracture set direction for the same cube of
Penrith sandstone of 13% porosity, plotted against
the permeability to flow parallel to the fracture set.
Results apply to the hydrostatic loading case, as in
Figure 13.56b. King et al., 1997.

In the subsequent hydrostatic loading, the velocities


tended to converge indicating near-closure, but the per-
meabilities of three similarly fractured sandstones,
(b) shown in Figure 13.58 reduced much more slowly with
stress increase than the smooth cleavage fractures in
Figure 13.56 a) Fracturing test cycle on a cube of Penrith sand-
the slate. The shear wave velocity Vs2 showed a contin-
stone of 13% porosity, with attenuation results for
uous rise.
the ‘3’ direction, perpendicular to the pending frac-
ture set, which presumably started to initiate at
King, 2002 reported ‘considerable hysteresis’ in Vs2 –
point F. b) Hydrostatic loading test phase for the permeability behaviour, with subsequent crack-closing
same set of fractures, with comparison to ‘before- cycles. An example is shown in Figure 13.58b. King was
fracturing’ phase. King et al., 1997. Note: seismic Q of the opinion that the reason for the significant differ-
values have been added on the right-hand axes, as the ences in velocity – permeability behaviour between the
geophysicists tradition for expressing attenuation as smooth and the rough fractures was ‘unclear’, but of
‘1000/Q’ is perhaps hiding a physically viable mecha- course cited the difference in roughness. For some reason,
nism related to attenuation, namely a certain modulus rough fractures closing due to stress increase contributed
increase due to normal loading, of unknown but simi- to increased velocity, but suffered a less-than-expected
lar magnitude to seismic Q, when the former is reduction in permeability.
expressed as GPa.
The reason for the different behaviour of the rough
fractures compared to the smooth, may be that E (phys-
fractures had a spacing of about 1 mm, while in the sand- ical aperture)  e (hydraulic aperture), for the case of
stone, fractured in the polyaxial cell, the rougher fractures rough fractures (or joints), and E  e for smooth frac-
were spaced many mm to 1–2 cm apart, based on a pho- tures (or joints). (Barton et al., 1985). This would mean
tograph of a fractured cube, provided in King, 2002. faster physical closure than hydraulic closure for rough
A particularly interesting, and possibly challenging joints.
result was obtained from the dynamically monitored per- The greater inequality with increased roughness JRC,
meability measurements. King and his colleagues had of the average physical aperture (E), compared to the
found that the low permeability (1 md) sandstone, theoretical, smooth-wall hydraulic aperture (e), described
started to develop measurable permeability once the more fully in Chapter 16, means that a fracture or joint
velocities Vp2 and Vs2 started to reduce (at point F in with an assumed (low) aspect ratio, say 104, may have
Figure 13.54), indicating crack development. a permeability equivalent to an even smaller aspect ratio
364 Rock quality, seismic velocity, attenuation and anisotropy

empirical equation of Barton et al., 1985 was as follows:

E2
e (13.16)
JRC 2.5

where JRC, the joint roughness coefficient of Barton


and Choubey, 1977, is fully explained in Chapter 16.
With E  100
m, and JRC  15, e  11.5
m. With
JRC  1, E  e  e.g. 10
m. Such a fracture offers no
resistance to closure: it closes to a significant degree under
even a very low normal stress.
Possibly the above result is one of the reasons that in a
(a) jointed or fractured reservoir, the performance of 4D seis-
mic can be quite successful if effective stresses are not too
high, as the velocity (and particularly the attenuation) are
relatively sensitive indicators of small permeability
changes. By the nature of jointed reservoirs, there are
unlikely to be commercially viable hydrocarbon-bearing
fractures or joints with very low JRC values, as joint clo-
sure under stress would preclude both permeability and
‘storage’, if such was needed due to low porosity matrix.
Hydraulic fractures, or widely opening joints, due to
the cooling and pressure-drive effect of water-flooding,
could be expected to show a ‘first-closing’ cycle show-
ing large hysteresis, when heat flow returns (see Bandis
(b)
et al., 1983 ambient tests).
Figure 13.58 Shear-wave monitoring of fracture closure and per- Velocity anisotropy due to a set of fractures, parallel or
meability reduction, due to increased normal load- perpendicular to principal loading axes, as beautifully
ing in the Imperial College D-H-M polyaxial cell. a) illustrated in the above polyaxial experiments is, inevi-
Three cubic samples of different sandstones, each tably, only a part of the information needed when trying
with a set of parallel, rough extension fractures, dis- to interpret field data from a jointed or fractured reser-
play quite different behaviour to smooth (cleavage- voir. Gibson and Toksöz, 1990 addressed the important
parallel) extension fractures in slate. b) Stress-closure question of crack orientations ranging from 0° to 90° in
cycles for the fractures in Crossland Hill sandstone,
relation to the stress axis. Cracks of small aspect ratio
show strong hysteresis on the first cycle. King, 2002
(roughly 104) were modelled, applying the theory of
and Shakeel, 1995.
Walsh, 1965 concerning crack closure under stress
applied obliquely to the crack normals. The variation of
the quasi compressional velocity with orientation, and
e.g. 105. This means that a 100
m physical aperture the important link to permeability variation with crack
may have a 10
m hydraulic aperture. orientation and stress level were each addressed.
So the physical aperture (E) of a rough joint or rough Some difficulty in closely matching the sparse exper-
fracture, with a typical JRC  10–15 for a well imental results was indicated, including matching the
controlled extension fracture, closes much faster than more rapid closure at low stress. The reality of fracture
the hydraulic aperture under stress, perhaps helping to sets having different properties affecting the stress-closure
explain the strong velocity response and the weaker per- behaviour, and the possible influence of intersecting or
meability response. When on the other hand the frac- interconnecting fractures also altering the stress-closure
ture has the lowest JRC of 1, typical for cleavage joints behaviour were discussed, but these difficulties could
in slate, there will be no inequality between E and e. not of course be modelled analytically.
JRC  1 is the practical limit to the following relation, In this complex rock mechanics area, there were by that
implying ‘table-top’ planarity and smoothness. The time promising advances in distinct element modelling
Rock physics at laboratory scale 365

Table 13.3 ‘Rule-of-thumb’ for order of magnitude estimates of seismic Qp, based on the empirical scheme
linking the jointed rock mass quality Q, Vp, Emass – and in very approximate terms – the rock mass
permeability. (Barton, 1999). Water at 20°C assumed in mD to m/s conversion

Vp km/s 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5
Qseis (est.) 2 3 5 7 10 15 22 32 46 68 100
K mD 1000 100 10 1 0.1 0.01
K m/s 105 106 107 108 109 1010

with the two-dimensional UDEC and three-dimensional There is something quite familiar about these esti-
3DEC codes of Cundall, allowing dynamic wave trans- mates of seismic Q in relation to Vp, based on the large
mission and attenuation through intersecting joint sets number of cases in fractured rock, reviewed in Chapter
of any desired orientation or block size. Various numer- 10, and in relevant parts of Chapter 13. Note that Vp2
ical modelling methods are discussed in Chapter 15. in Figure 13.54a is 2.5 km/s at the end of the cracking
phase, just as the reported seismic Q of a second sand-
13.14 Rock mass quality, stone was 5 at the same stage of loading, following the
attenuation and modulus same method of fracture development.
Beyond the ‘jointed rock’ modulus limit of roughly
This chapter will be concluded by bringing an empiri- 150 GPa, the question of seismic Q value estimation is
cal rock mechanics scheme into consideration, as justi- problematic, but remarkably it starts to resemble the rock
fication for the tentative, repeated conclusion, first mass Q and Qc value, i.e. 500, 1000, 2000 – the latter
launched in Chapter 10, that seismic Q very much representing completely intact hard rock at moderate,
resembles the static modulus of deformation expressed deep and kilometre-depths, or relevant to hard, very hard
in GPa, strictly for the case of jointed or fractured rock, for and extremely hard unjointed rock at moderate depths
which a good data base exists. The special fracturing of say 1⁄2 to 1 km.
tests of King and his colleagues at Imperial College, In case of knowledge of the approximate uniaxial
have given repeated support for this simple hypothesis. compressive strength ( c) of the rock where the P-wave
Deformation moduli are almost always in the range of velocity measurement is made, the improved linkages
5 to 100 GPa, and seismic Q seems also to be most fre- between ( c) and deformation modulus (and therefore
quently in this range in the case of near-surface rock an approximate seismic Q ) can be applied, as shown in
masses (to 1 kilometre depth?), as frequently suggested Figure 13.60. Example: Vp measured at 2 km depth in
by attenuation data presented in Chapters 10 and 13. fractured sandstones  5.0 km/s. Uniaxial strength of
The empirical expression for static deformation modu- sandstones 50 MPa. Estimate of ‘static’ deformation
lus: Emass  10 Qc1/3, and the expression linking vel- modulus is 25 GPa. A first estimate of seismic Q,
ocity and rock quality: Vp  3.5  log Qc can be specifically Qp is therefore 25.
merged, by elimination of Qc into the form: The possible relative differences in frequency effects on
Vp and seismic Q is of course a source of additional error
in this simple method. Porosity effects not captured by
E mass  10(Vp0.5)/3 (GPa ) ( Q seis ?) (13.17) direct effects on ( c) may be subtracted from the esti-
mates of modulus, using the graphical adjustments sug-
The Vp scale on the left-hand side of Figure 13.59 can gested in Figure 13.59. These are empirical by nature and
be followed across the diagram (‘ignoring’ the rock mass have an insignificant data-base at high rock quality Q-
quality Q c), all the way to the right-hand side deformation values, since high n% seldom accompanies high Q-values.
modulus estimates, giving (in Table 13.3) proposed Vp – In Chapter 7 concerning excavation disturbed zones,
Emass (GPa) or seismic Q ‘first-pass’ estimation of (better we reviewed the very thorough seismic studies of the
than) order of magnitude values of seismic Q, specifically Basalt Waste Isolation Project (BWIP) conducted by
where jointed or fractured rock masses are involved. For King et al., 1984 and 1986, and several other researchers.
good measure, a very rough estimate of permeability is The review included EDZ effects on P- and S-wave
also given, based on the Lugeon value (1 L  107 m/s, velocities in the columnar basalt, using cross-hole seismic
and with 1 Darcy  1012 m2  105 m/s, for water at measurement at different depths into the tunnel wall in
20)C, we have 1 Lugeon  10 millidarcies). horizontal, diagonal (inclined), and vertical directions.
366 Rock quality, seismic velocity, attenuation and anisotropy

Figure 13.59 For the case of jointed or fractured rock, there is a strong resemblance of the magnitude of seismic Q, to the ‘static’ deformation
modulus expressed as GPa. The figure shows the inter-relationships developed from an empirical expression linking the rock
mass quality Q-value with the ‘static’ deformation modulus Emass (or M) and velocity Vp. Permeability, in Lugeons
(1 L  107 m/s  10 millidarcies for water at 20)C) is also roughly linked, in the absence of clay-sealing of the joints. In this
chart a nominal uniaxial compressive strength of ‘hard’ rock, namely 100 MPa forms the basis of the graphics, but using the
Q c  Q  c/100 relation, other compressive strengths can be applied, as shown in Figure 13.60. Barton, 1999a.

Figure 13.60 A rough method for estimating seismic Q from P-wave velocity and uniaxial compressive strength linkages to ‘static’ deforma-
tion modulus. By implication, and also in practice, the trend when moving downwards-and-to-the-right is for increased depth
and reduced porosity, thereby reaching values of Emass in the higher range of 50 to 100 or more typical for crystalline rocks or
hard limestones or well-cemented sandstones.

A very interesting result of relevance to the possible Q p value in situ cannot be more than that of an intact
link between seismic Q p and the static deformation mod- sample of the same rock at the same stress level. King et al.,
ulus determined from Q rock is shown in Figure 13.61. 1986 normalized their in situ estimates of Q p, using
This is not laboratory data for Q p, but it is normalized an assigned maximum value of 50, where the highest
by laboratory data, on the assumption that the relative value of Q seis was recorded, 9 m from the tunnel face, in
Rock physics at laboratory scale 367

a vertical direction (not crossing any columnar joints,


and probably few if any horizontal cross-column joints).
The value of 50 corresponded to the laboratory test
value at appropriate low stress levels.
Note the horizontal ray-path relative Q of about 5 to
8, measured in a direction crossing the maximum num-
ber of joints (sinuous six-sided basalt columns of 0.15
to 0.36 m thickness), and possibly under the influence
of a low horizontal stress, typical of many near-surface
columnar basalts. The depth of the tunnel was only
46 m. In the vertical direction, with theoretical vertical
tangential stress effects perhaps more clearly imprinted,
due to the relative lack of horizontal joints, there is an
unmistakable similarity of the increase in relative Q to
the deformation modulus increase (and magnitude,
when expressed as GPa) that one often interprets from
tunnel and shaft extensometer arrays.
For example Barton and Bakhtar, 1983, back calcu-
Figure 13.61 Relative Qp values interpreted from seismic cross- lated static deformation moduli from extensometer
hole measurements in the face of a shallow (46 m) measurements in a deep (1.6 km) shaft in steeply bed-
tunnel. Measurements in horizontal (strongly jointed) ded and jointed quartzites, obtaining moduli from 2.5
and vertical (sparsely jointed) directions, in thin- to 60 GPa from near the shaft wall to one diameter
column basalts of 0.15 to 0.36 m thickness at BWIP. depth. There is also an unmistakable similarity of these
King et al., 1986. numbers to expected Qp ranges, as also in Figure 13.60.
14 P-waves for characterising
fractured reservoirs

The ability to detect the presence of viable hydrocarbon- velocity-depth function attributed to Slotnick, 1936:
bearing structures, with sufficient porosity, and with
tolerable matrix and mass permeability, are among the (14.1)
v z  v o  Kz
challenges of the petroleum geologist, petrophysicist,
and geophysicist, whose joint role may last far into the
reservoir engineers’ production phase. In this chapter where vo is the (P-wave) velocity at the surface and vz is
we will give examples of some of the basic ways of ‘seis- the velocity at vertical depth z.
mically illuminating’ reservoirs at larger scale, some- There was also early recognition of a quite systematic
times extending over 100’s of km2 to depths of 5 km. trend linking velocity to the geological age, in combin-
(Small-scale dipole and monopole sonic logging was ation with the present depth of occurrence. The interest-
described in Chapter 12, because of its intimate con- ing empirical method of Faust, 1952 will be summarized
nection to rock stress, borehole stability and mud pres- as a ‘geological’ and ‘stress-effect’ introduction to hydro-
sure). The larger scale techniques illustrated in this carbon reservoir investigation.
chapter range from cross-well tomography, VSP 2D Faust discovered that the greatest rate of velocity
and 3D multi-azimuth walk-away surveys, 4C multi- increase occurred at shallow depth in the oldest units, which
component surveys, AVO and AVOA for detecting is fundamental early proof of the importance of dual
fracture orientation, and 4D repeated surveys for track- porosity. The likelihood of more joints in the stiffer, older
ing reservoir changes over time. One of the causes of units make these units more sensitive to stress change.
such changes is the use of water flooding, which causes However, with only Vp as a dynamic indicator of condi-
various coupled mechanisms, besides an advancing tions, acoustic closure represented a limit to the sensitiv-
oil/water contact. 4D can also be used for monitoring ity, especially for the weaker, younger reservoir rocks.
the effects of compaction, and subsidence. In general, Faust, 1951, used data from almost 1 million feet (or
apart from a brief treatment of C-waves or converted about 300 kilometres) of well sections, in 500 petrol-
P-S waves, we will leave a detailed description of the use eum well surveys, mostly from the USA and some from
of shear wave splitting and polarization for Chapter 15. Canada. The great majority of data was for mixed
This remarkable method for characterization of frac- shale/sandstone sections. A non-systematic comparison
tured reservoirs is a suitable finale, and a good intro- of shale and sand (sandstone) velocities had revealed an
duction to the need for more geomechanics (in Chapter average discrepancy of only 350 ft/sec, or 106.7 m/s in
16), for improved understanding of fractured reservoir velocity between these two, frequently inter-bedded
behaviour. units, the sandstone having the highest velocity by this
small average margin.
The ironic similarity of velocities for these two basic
14.1 Some classic relationships dissimilar as ‘chalk-and-cheese’ units, is a reminder of the
between age, depth and potential ‘non-uniqueness’ of P-wave velocity, and the
velocity recognised need for alternative interpretation methods,
such as impedance and attenuation, to distinguish the
The most fundamental and earliest means of interpret- different lithologies and their fluid-bearing signatures.
ing possible reservoir conditions at kilometre depths was The potential closeness of velocities for shale and
the use of seismic P-wave velocity well surveys, using a sandstone is actually surprising, in view of the greater
simple VSP concept. There was early recognition of tolerance of the stronger sandstone to stress anisotropy,
velocity increase with depth, with the following smooth often resulting in several MPa greater minimum stress
370 Rock quality, seismic velocity, attenuation and anisotropy

Table 14.1 Mean depth-velocity data from Tertiary and of velocity increase occurs at ‘shallow’ depth in the oldest
Pennsylvanian shale-sandstone units, selected from units, which is presumably a function of more joints that
Faust, 1951, from wells in the USA only. therefore make the older units more sensitive to stress
Mean Mean change. The steepest velocity-depth gradients occur, of
Mean velocity No. of velocity of No. of course, in the 300 m to 1 km depth range, 1000 ft being
depth of Tertiary wells of Pennsylvanian wells of the approximate limiting depth of measurement.
Z (ft) (ft/s) Tertiary (ft/s) Pennsylvanian The trend of the data is given by Faust, 1951, as:
1 025 6 800 71 9 420 18 1
V  m Z n (14.2)
2 500 7 660 63 11 110 14
3 500 8 160 63 11 720 18
4 500 8 670 64 12 230 18
which was reportedly similar to a current weathering
5 500 9 220 54 12 650 14 correction method known as the ‘Blondeau weathering’.
6 500 9 520 53 12 710 7 Z is depth in feet, m is a constant with units of velocity
7 500 9 860 48 13 320 5 (that proves to be age-dependent) and n is a constant,
8 500 10 220 31 13 390 4 independent of age.
9 500 10 670 23 13 020 4 Faust plotted the same data in log-log format, as
10 500 11 090 13 14 030 3 shown in Figure 14.2. The velocity at each depth could
11 500 11 300 2 14 500 2 then be represented as:
log V  log m  1 n log Z (14.3)
in the shale, which is frequently a fluid barrier for the
hydrocarbon-bearing sandstone. To the extent that equation 14.2 is correct, m
Faust minimised lithological variations by averaging all represents the velocity for a particular geological age
measurements at the same depth and age. At the outset he at 1 foot depth (0.3 m). The relevant geological ages
assumed that velocity (V) could be expressed as a func- and the ‘zero depth’ velocity constants (m) are given in
tion of depth (Z), age (T) and lithological variables (L). Table 14.2. The constant n  6, applies for all the
Since L was considered too problematic (i.e. how to com- curves in Figure 14.2.
pare the separate units limestone, shale and sandstone) By plotting m against age, shown in Figure 14.3, a
shale and sand were accepted as representing equivalent linear log-log plot was obtained of the form:
sections, as they alternated too frequently in relation to
the usual interval of down-hole velocity measurement, am  T1/6 (14.4)
which then was as much as 500 feet (or 152 m). Table 14.1
gives an extract of the data for the two most frequently where T is geologic age in years and  is a constant.
represented geologic ages, namely the Tertiary and the Equations 14.2 and 14.4 could therefore be combined as:
Pennsylvanian, which follows the Permian. The oldest
V  (TZ)1/6 (14.5)
units were Devonian (15 wells) and Ordovician (3 wells).
Although reproducing very ‘smoothed’ data due to the In Faust, 1951,  is given as 125.3, and is numeric-
averaging process, Faust, 1951, was careful to point out ally equal to velocity in feet per second when TZ  1.
that the maximum deviations from the velocity averages From the literature of the time, values of  of 121.5
could be great. Nevertheless both minimum and max- and 127.2 were also quoted by Faust. A mean value of
imum values also demonstrated increased velocity with 124.7 from these three results suggests that   125
depth. Comparison of same-age same-depth data from would be an acceptable constant.
different regions of the USA, reportedly showed little sys- An example will now be given to illustrate this simple,
tematic deviation of the mean from one area to another. early technique.
However the Devonian of the Appalachian Basin and the
Eocene and Cretaceous of SW Texas had velocities more
typical of data 3000 ft (915 m) deeper. This was thought Assume Pennsylvanian  220 106 years (Table 14.2)
to be due to unusually high degrees of cementation. M  3,047 ft/s  929 m/s (Table 14.2)
Evaluating Z  1,000; 1,500; 2,000; 3,000 ft, we obtain:
Figure 14.1 is derived from the extensive data that was
equation 14.5 at V  9,712; 10,391; 10,900;
partly sampled in Table 14.1. It shows the quite systematic
11,663 ft/s  2.96; 3.17; 3.32; 3.56 km/s
average trends of velocity, age and depth. The greatest rate
P-waves for characterising fractured reservoirs 371

Figure 14.1 Velocity-depth trends for in situ reservoir rocks of various geological ages (linear scales). Faust, 1951.

Figure 14.2 Velocity-depth trends for in situ reservoir rocks of


various geological ages (log scales). Faust, 1951. Figure 14.3 Mean values of m (Velocity in ft/s at nominal 1 ft
depth) for different ages. Faust, 1951.
Table 14.2 Geologic age and ‘zero depth’ velocity constants.
Faust, 1951.

Velocity Velocity These predicted velocities at 305 m, 457 m, 610 m


constant constant and 915 m depth, are apparently much less sensitive to
Geologic age Geologic time m ft/s m m/s stress increase than jointed rock would be to sub-hori-
Tertiary 26 106 yrs 2 190 668 zontal wave paths, from refraction seismic or cross-hole
Eocene 43 106 yrs 2 332 711 tomography. Perhaps this is because we are looking at an
Cretaceous 93 106 yrs 2 607 795 inter-bedded sandstone-shale unit of Pennsylvanian age,
Jurassic-Triassic 152 106 yrs 2 823 861 with the predominant measurement direction in a sub-
Permian 192 106 yrs 2 866 874 vertical direction, with shot points only some 1,000 ft
Pennsylvanian 220 106 yrs 3 047 929 (305 m) from the well, and with downhole receivers to
Mississippian 245 106 yrs 3 235 986 record the approximately 0.09 to 0.08 s average arrival
Devonian 284 106 yrs 3 380 1 030
time, limited to each 1,000 ft depth interval (in the case
Ordovician 390 106 yrs 3 439 1 048
of the Pennsylvanian age sediments).
372 Rock quality, seismic velocity, attenuation and anisotropy

from 4% to 39% porosity, with clay volume fractions


spanning 0 to 55%, and over a confining pressure range
of 0 to 40 MPa, equivalent to depths up to roughly 3 km.

14.2 Anisotropy and heterogeneity


caused by inter-bedded strata
and jointing

Fine layering of alternating porous and impermeable


strata is obviously one of the basic sedimentary systems
that contribute to the existence of potential reservoir
rocks in sedimentary basins. Sandstones with their usu-
ally elevated porosities, may be the recipients of the
hydrocarbons, and the variously aged shale facies may
contribute both as the deeper and or laterally occurring
source rocks, and as the elevated seals or cap-rocks. Some
form of deformation of the recipient strata, through fold-
ing or faulting, is usually necessary to ensure that a trap
is formed. Petroleum may therefore be found within
anticlinal structures containing favourable jointing, if
capped by impermeable shales or salt rock. Frequently it
is the larger scale faulting that is responsible for the
porous reservoir rocks and sealing layers to be juxta-
posed, or sealed by shale-smear if sufficiently plastic. The
Figure 14.4 Remarkable similarities of Vs versus Vp trends for final requirement is a favourably tilted dip direction.
water saturated sandstones and shales, according to Fine layering of sedimentary strata means that the
ultrasonic testing in the laboratory at effective dominant wavelength of a seismic or sonic pulse is long
pressures of up to 40 MPa. Castagna et al., 1993, as compared to the thickness of individual layers. The
reproduced in Mavko et al., 1998. medium will nevertheless exhibit effective (and real)
anisotropy, with a vertical symmetry axis in the case of
horizontal layering. In the presence of hydrocarbons this
layered medium may show substantial attenuation and
The natural suspicion that the ‘coarse’ VSP-style well velocity dispersion, which will be compounded with the
logging available at this time may have been the cause additional (or separate) presence of jointing or fracturing.
of the poor differentiation of the shale and sandstone in The authors Helbig and Thomsen, 2005 emphasised
terms of their relative P-wave velocities, as referred to at that anisotropy, and the associated new techniques:
the beginning of this review of Faust, 1951, is defini- primarily shear wave splitting and polarization, bought
tively not supported by confined ultrasonic laboratory with them a new exploration concept: neither exploring
data for shales and sandstones. Figure 14.4, from work for the presence of reservoir rock, nor for the presence of
by Castagna et al., 1993, and reproduced by Mavko hydrocarbons, but for the presence of crack or fracture
et al., 1998, shows that the three given empirical rela- permeability. In 4D repeated surveys, changes of fracture
tionships for Vp and Vs seem to fit both data sets very permeability due to flooding and production has also
closely. Han’s easy to remember 1986 relationship: become a modern goal in the use of seismic and attenu-
ation anisotropy.
Vs  0.79Vp  0.79 (km/s) (14.6) Following Lynn, 2004, a ‘thin bed’ is 3 8 of a wave
length, the limit for a discrete reflection both from the
top and bottom of the bed. Wave scattering, attenu-
was shown by Mavko et al., 1998 to give a very good ation and dispersion occur when the ordered hetero-
fit to a wide variety of (water saturated) shaly sands, geneities have scale lengths of about 0.3–0.01 of the
P-waves for characterising fractured reservoirs 373

wavelength, while the smallest scale of ordered hetero- 4. Type IV: macro-porous reservoirs have high matrix
geneity, less than 0.01 of the wavelengths, may be the porosity and permeability – matrix provides both
cause of most of the azimuthal and offset dependent vel- storage capacity and fluid-flow pathways – frac-
ocity. Conventional seismic wavelengths are much longer tures merely enhance permeability.
than the scale lengths of either of the features that gov-
ern dual-porosity flow in a reservoir. The author warned that Type I and Type II reservoirs
As pointed out by Williams and Jenner, 2002, the could be easily damaged by excessive production rates
earth does not care which tools or frequencies we use; it (due presumably to the rather strong sensitivity of per-
still knows it is anisotropic. Strong P-wave velocity meability to unnecessarily high effective stress levels),
anisotropy is being observed in every geologic environ- but many performed well under unassisted primary
ment, with the possible exception of basins under pri- recovery when managed properly. In Type III reservoirs
mary deposition and burial. P-wave azimuthal anisotropy the recovery factor was dependent on lithology,
was previously ignored, and left to the research and wetability, and fracture intensity. The choice of appro-
technology group, but is now known to be one of the priate EOR (enhanced oil recovery) was essential
most significant properties of the acquired seismic data. for optimal exploitation. In Type IV reservoirs, the
Unfortunately, in the marine environment, fully popu- recovery factor was most sensitive to the selected drive
lated offsets in each azimuth class are less common than mechanism.
on land, but Williams and Jenner emphasised that even The sophistication of investigation methods, using
narrow azimuth data gave an opportunity to see the multi-component and multiple-frequency methods, gives
effects of azimuthal anisotropy. the capability of revealing heterogeneity and fracturing
Hand-in-hand with the basic anisotropy caused by at many scales, even if indirectly, thanks to some remark-
sedimentary layering, and deformation processes, is able and fortuitous dynamic wave properties.
marked heterogeneity, occurring at many scales and for A good analogy to the developing heterogeneity of a
many reasons. As Nur, 1989 pointed out, reservoirs are better understood reservoir, is all the adverse faulting
much more heterogeneous than anybody likes to believe, gradually revealed by successive drilling at a potential
and as time goes by more and more reservoirs are nuclear waste disposal site, perhaps causing its eventual
re-classified as severely heterogeneous, due to a multitude rejection, after many years of costly investigations.
of dynamic flow-related cyclical events during their for- With huge quantities of petroleum in place at a poten-
mation, and due to fracturing and faulting in subsequent tial or existing reservoir, rejection is seldom an option,
geologic eras, each of which become better understood as and better understanding through improved seismic
time, and seismic developments, advance. An appropriate and enhanced production techniques, are the obvious
quotation from Lynn, 2004, also has relevance here. ways forward.
‘Fractures are like cockroaches. There is no such thing as
one cockroach. If you see one, a whole family of all scale
lengths is hiding nearby.’ 14.2.1 Some basic anisotropy
In a survey of one hundred fractured (i.e. jointed, not theory
MHF) oil reservoirs from around the world, it was found
convenient to divide the reservoirs into four groups For reference purposes we need to summarize some basic
(Allan, 2002): elements of isotropic and anisotropic behaviour, since
various categories of anisotropic behaviour will now be
1. Type I: little matrix porosity and permeability – treated in somewhat more detail than in earlier chapters.
fractures provided both storage capacity and fluid- A linear elastic isotropic medium requires only two
flow pathways. constants to specify the stress-strain behaviour, either
2. Type II: low matrix porosity and permeability – Young’s modulus (E) and Poisson’s ratio (*), or alterna-
matrix provides some storage capacity – fractures tively Lamé’s constants ( ) and (
), where (
) is the
provide fluid-flow pathways. shear modulus. These pairs of parameters can be
3. Type III: micro-porous reservoirs with high matrix derived from each other using standard equations of
porosity but low matrix permeability – matrix pro- elasticity (see Chapter 1, and refer also to the Rock
vides the storage capacity – fractures provide the Physics Handbook by Mavko et al., 1998, and Birch,
fluid-flow pathways. 1961).
374 Rock quality, seismic velocity, attenuation and anisotropy

Matrix format (6  6) is used to represent the three- ● Transversely isotropic with horizontal axis of sym-
dimensional elastic tensors, which in the isotropic case, metry (TIH) perpendicular to the vertical layers, but
requires only two parameters, as follows: typically caused by stress-related aligned, vertical
jointing, fracturing, and/or microcracks. (Also five
independent constants: tensor elements 1,2 and 1,3
  2
0 0 0  are the same term c13)

  2
0 0 0  Orthorhombic symmetry typical of horizontally
0 

  2
0 0

0 0  (14.7) bedded rock containing a set of vertically aligned
  joints, fractures and/or microcracks. (Nine inde-


pendent constants).

Symmetry, as above, is an important property, if found,


as the wave fronts will be similarly symmetric. When pro-
where the term  2
can be expressed as c11, as c12, cessing orthorhombic data (which can have a second per-
and
as c44. pendicular fracture set and still be termed orthorhombic),
Figure 14.5 shows the contrasting elastic tensors for if fast and slow directions have been identified, then azimuth
three classes of anisotropy: sectoring can be applied in these directions, and according
to Lynn, 2004, most ‘isotropic’ processing codes can func-
● Transversely isotropic with vertical axis of symmetry tion fairly well on these azimuth-sectored data.
(TIV) typical of fine layering in shales. (Five inde-
pendent constants, since c66  1⁄2(c11 – c12). Tensor
elements 1,3 and 2,3 are the same term c13) 14.3 Shallow cross-well seismic
tomography

For consistency of presentation we may start this brief


review of reservoir characterization, by referring to an
interesting shallow 3D tomographic imaging of both
seismic velocity and attenuation, given by Brzostowski
and McMechan, 1992. This was based on a 1989 3D
survey shot by Peko Oil USA, and by Texaco in SE
Oklahoma. Figure 14.6 (a to d) shows the surface geol-
ogy and Vp tomograms for the first, second and third
constant-thickness slices centred at depths of 32, 96 and
106 m. The corresponding seismic Qp tomograms are
shown in Figure 14.7.
As is readily noted from this interesting set of data,
the alluvium (e.g. unit Qal) had both low velocity
(2.25 km/s) and low Qp (6) at 32 m average depth,
while the lithified sandstones had velocities between
2.5 and 3.5 km/s at 32 m average depth and Qp values
of about 7 to 10. Clearly the Vp values for all units and the
corresponding Qp values increase with depth. However,
more lateral variation is shown in the Qp tomogram. As
we have seen both in Chapter 1 and in Chapter 11, this
Figure 14.5 Elastic tensors in 6  6 matrix format, for two single lateral variability is also common at much larger scales.
‘geological’ classes of symmetry (TIV and TIH), com- The suggested link between Vp and Qp developed at
bined as more realistic orthorhombic material in the
the end of Chapter 13 (Figure 13.60) suggests Qp val-
third diagram, and requiring nine instead of five inde-
ues lower than 2 to 3, in the case of weak materials in
pendent constants. Typical of bedded or layered rock
with anisotropic horizontal-stress related, aligned
the presence of velocities lower than 2.25 km/s. Where
vertical jointing, fracturing and/or microcracking. velocities climb to 3.5 km/s in much of layer 3 (centred
P-waves for characterising fractured reservoirs 375

Figure 14.6 3D P-wave velocity tomograms for mid-depths of 32, 96 and 106 m depth in quaternary sediments, alluvial stream beds and
sandstones. Brzostowski and McMechan, 1992.

at 106 m depth), the measured Qp was predominantly


from 5 to 10, and Figure 13.60 suggests Qp of at least
5 to 7, in the case of moderately weak materials, such as
‘shallow’ sandstones.
Since both Vp and seismic Q contain a certain level
of information about the pore space and degree of joint-
ing or fracturing, and the saturating fluid, the trial
combination of these parameters in the form of tomo-
graphic plots of Qseis/Vp, possibly offers the possibility
of a closer estimate of transport properties.
In an interesting study using tomographic imaging
between a total of six wells with known lithology, but
unknown transport properties, Liao and McMechan,
1997 showed what might be delineated by the combin-
ation of tomographic images of Vp, seismic Q, and
their ratio Qseis/Vp. Figure 14.8 shows the 210–400 m
deep shale/sandstone sequence. The Gypsy sands were
high porosity (mean 20%), high permeability (mean
560 mD), water-saturated, clean sand channels lying
within a silty flood-plain.
High-resolution cylindrical piezoelectric bender
transducers were used as both sources and receivers,
over a band width of 300–1300 Hz. From the diamond-
pattern of six wells, the data between wells 1, 5 and 7
were selected for Qseis and Vp tomographic analysis.
Figure 14.7 3D Qp tomograms corresponding to Fig. 14.6 slices. Overall full-depth, three-hole, results are shown in Figure
Brzostowski and McMechan, 1992. 14.9, while a detailed 90 m section between wells 7 and
376 Rock quality, seismic velocity, attenuation and anisotropy

Figure 14.8 Sand, shale and sand-shale lithology at the three-well


tomographic site described by Liao and McMechan,
1997.

5 is shown in Figure 14.10. This has a marked shale-rich


upper 40%, and a sand-rich middle-to-lower section.
The typical in situ values of the parameters were as
follows:
Vp Qseis Qseis/Vp
sandstone 3.0 km/s 45 15
shale 3.6 km/s 30 8.3

The low velocity was a consequence of the high


porosity (mean 20%) of the channel sand. The relatively Figure 14.9 Fence diagrams of seismic tomography analysis between
low attenuation (Qseis as high as 45) was reportedly due wells 1, 5 and 7 to full depth. The wells were 99 m,
to its full water saturation, and high porosity. The 141 m and 105 m apart. Liao and McMechan, 1997.
authors considered it likely that high values of the ratio Note Vp, Qseis and Qseis/Vp tomograms. Liao and
McMechan, 1997. Reproduced by kind permission.
Qseis /Vp would correlate with the most porous and most
permeable zones, and low values with reduced flow such as them less attenuating (no squirt flow losses), giving a
the shale-rich layers. They considered that a thin and a relatively high Qseis of 45. This combination (ratio 
thick shale barrier did show appropriately low values of 15) is many times higher than the modest ratios of Qp /Vp
the ratio (i.e. light colours in Figure 14.10, panel c). suggested in the jointed rock model suggested in Figure
The authors cited Castagne et al., 1993, who had 13.60, at the end of Chapter 13.
suggested that sandstones often have higher velocity Perhaps, on reflection, this contrast in Qseis/Vp ratios,
and show less attenuation than shales in the same envir- being so marked, could delineate the difference
onment. In these channel sands the low velocity between unconsolidated sands and jointed sandstones,
(3 km/s) is clearly a function of lack of cementation and between weak plastic shales, and the less desirable
and their high porosity, while the full-saturation makes fissured/jointed, or indurated variety.
P-waves for characterising fractured reservoirs 377

shallow (15 to 35 m) water saturated limestones, was to


image gas filled fractures in wells with 50 to 100 m sep-
aration. The authors had started out using standard low
resolution VSP and surface reflection methods, which
yielded little information on the individual fractures,
but did indicate P-wave anisotropy due to a N70°E
trending set of joints.
Cross-well, high frequency (1 to 10 kHz) surveys using
piezoelectric sources, were later used with success, to
image an individual fracture zone in the limestone,
before, during, and after air injection. The air followed
the hydrologically-predicted pathway, and helped to
seismically ‘illuminate’ the zone. Subsequent slant-hole
coring verified the presence of a conducting, vertical
fracture 1 m from the seismically located position.
The wave-lengths used were of the order of fi to 1 m,
and the fracture aperture reportedly as large as 1 mm.
The P-wave velocity of the limestones determined
from earlier near-offset VSP varied from about 3.8 to
4.2 km/s between the shallow depths of 16 and 26 m,
which suggests a rock quality Q-value of 2 to 5 from the
empirical relation Vp  3.5  log Q derived in Part I,
relevant to nominal 25 m depth, ‘100 MPa, low-porosity’
rock. Q  2–5 is typical for rock masses with three sets
of joints, moderate block size, and with possible wea-
thering of the joint walls: i.e. from Appendix A: Q 
90/9  2/2  0.66/2.5  2 to 3.
A seismic Q of the order of 5–20 times higher than
Qrock, as suggested from numerous cases reviewed in
Chapter 10, suggested an ‘order of magnitude’ estimate
of Qseis  10–25, at this shallow, jointed site.
The method suggested in Figure 13.60, and Table
13.3, linking Qseis to the ‘static’ deformation modulus
estimate, suggests a Qseis value of about 13–14 from
Qrock  2 to 3, if the uniaxial compressive strength of
Figure 14.10 Selected tomographic images of Vp, Qseis, and their the limestone was around 100 MPa. Alternatively, util-
ratio Qseis/Vp, between well 7 (left) and well 5 (right), ising Figure 13.60, for evaluating the possible effect of
for contrasting shale-rich (top), and sand-rich (lower- lower uniaxial strengths, we see that Qseis for this jointed
middle) zones within a 90 m section between the (i.e. fractured) limestone might be about 12 to 16 with
wells. When dominated by shale, ratios of Qseis/Vp are
the given Vp range of 3.8–4.2 km/s.
low, and when dominated by sand they are high.
Sand/shale proportions for each borehole are shown.
Liao and McMechan, 1997. Reproduced by kind
permission. 14.3.2 Cross-well seismic
tomography with
14.3.1 Shallow cross-well seismic permeability measurement
in fractured rock
A new poroelastic analytical model called ‘super – k’ was
The question of resolution levels necessary for fracture described by Yamamoto, 2003, who was working not in
detection was illustrated by Majer et al., 1997. Working petroleum provinces, but on a limestone aquifer. Pointing
at the Conoco Borehole Test Facility, their goal in out that extraction of permeability from velocity and
378 Rock quality, seismic velocity, attenuation and anisotropy

Table 14.3 Seismically derived, frequency-dependent cross-well


data for Vp and Q, with the seismically derived (super-k)
prediction of permeability and porosity. Limestone
section at 462 m depth. Yamamoto, 2003.

Frequency (kHz) Vp (m/s) Q k (d) 

2 3545 14.0 33.7 0.34


4 3580 18.1 33.7 0.35
8 3616 31.0 35.0 0.36
12 3652 49.0 35.7 0.34

Note: k from pumping test  33.2 d


k from packer test  36.3 d
porosity from neutron log:   39%

attenuation data had been one of the important unsolved Figure 14.11 Comparison of models (K  super – k, B  Biot,
and D  BISQ ) with cross-well velocity and attenu-
problems in seismic exploration, the author was able to
ation data from 300–480 m depth at a permeable
demonstrate good correspondence between acoustically-
aquifer site. Data from Yamamoto and Kuru, 1997.
imaged permeability and hydraulic pumping (extraction) Yamamoto, 2003.
and packer (injection) tests. An unusual close, 11 m
horizontally-spaced cross-well set-up was used for deriv-
ing 4 kHz velocity and attenuation tomograms. Data was
also acquired at five other frequencies (1, 2, 3, 5, and it would presumably resemble chalk more than lime-
0.250 kHz). Permeability was measured in four packered stone). Significantly, the mismatch with the Qseis modu-
intervals down the same 300–480 m deep test section. lus-model (Figure 13.60) increases strongly at higher
His model reportedly coincided numerically with the frequencies.
combined Biot and squirt flow mechanism (the BISQ
model of Dvorkin et al., 1994), when permeability
was 100 md, and when the frequency was 100 kHz 14.3.3 Cross-well seismic in deeper
(the so-called super-k regime, where the pore fluid is reservoir characterization
always relaxed). The limestone where the author per-
formed these studies had extremely high permeability A special issue of Geophysics was devoted to cross-well
(10–250 d). He described a large wave attenuation and methods in 1995, due to the large number of unsoli-
velocity dispersion in this highly permeable limestone. cited papers on the subject, following strong oil indus-
This in situ data is reproduced in Figure 14.11. try interest in use of cross-hole tomographic methods
The author divided the 180 m  11 m imaged cross- since the late 80’s, for imaging below the resolution of
section into seven, mostly 25 m long sub-sections, for surface seismic, for such purposes as steam injection
the purpose of velocity and attenuation inversion. His front imaging. An important discovery from the early
analytical inversion model for permeability involved a period of cross-well research, was that very high fre-
quadratic equation for a non-dimensional permeability- quency could be propagated over distance of several hun-
frequency parameter, with attenuation as a polynomial dreds of metres, when both source and receiver were
constant. down-hole, in deep boreholes.
It may be noted that although the porosity and per- The additional discovery of very high bandwidth
meability were of exceptional magnitude, the seismic with downhole receivers and sources apparently con-
velocities of 3.5 and 3.6 km/s obtained at the two lowest firmed the phenomenon that seismologists also have to
frequencies of 2 and 4 kHz, were matched with seismic live with, if without downhole instruments (Chapter
Q values of 14 and 18, which are quite close the jointed 10), that most of the surface-based frequency loss is due
rock model potentially linking Vp and Qp, if the assump- to attenuation in the near-surface low-Q zone. Rector,
tion is made that the uniaxial strength of the limestone 1995, mentioned an order of magnitude improvement
is high, and that the high porosity and permeability is in subsurface resolution, with many investigators now
due to solution channels. (If porosity was distributed using secondary arrivals, such as reflections, to obtain
P-waves for characterising fractured reservoirs 379

very high resolution images. As at the surface, where


refraction imaging was gradually replaced by reflection
imaging, the same occurred with VSP, and now was
occurring when receiver and source were downhole.
Mathisen et al., 1995 gave an example of the use of
time-lapse cross-well data in heavy oil sands, using a
cemented multiple-receiver cable in one hole. Twenty
seven cross-well surveys were acquired between two wells,
during a 31⁄2 month period, before, during, and after a 34
days steam injection cycle. Vp and Poisson’s ratio tomo-
grams were used to track the changes caused by both the
temperature and viscosity reduction. Stratification with
a dip of 20° controlled the flow of steam, which caused
progressive reductions in P-wave velocity of up to 90 to
270 m/s adjacent to and above the injectors. Poisson’s
ratio tomograms showed a corresponding decrease (0.1),
in the same areas. Time-lapse S-wave tomograms demon-
strated great stability with time, showing only very slight
change (i.e. shear waves not registering fluid viscosity
changes), and therefore demonstrating that the reservoir
‘rock’ was stable despite the fluid changes.

14.4 Detecting finely inter-layered


sequences

Remaining at shallow depth for the moment, namely a


250 m deep research well through finely inter-layered
limestone, shale, and sandstone sequences in the Imperial Figure 14.12 a) The results of multiple borehole logging tools in
College Borehole Test Site, one may note the important resolving the velocities and densities of finely inter-
work by Sams, 1995, who showed how the combined use layered rocks. b) The relative velocities of the shales,
of a borehole compensated (BHC) sonic logging tool, a sandstones and limestones from laboratory tests of
compensated formation density tool, and a Formation Vp, and an assumed Vs correlation. Sams, 1995.
Micro Scanner (FM) could resolve much of the detail of
finely interlayered rock sequences.
Figure 14.12 shows the success of multiple borehole important for remote detection of fluids, such as kerogen
logging tools in resolving some details of finely inter- rich shales and their state of maturation.
layered rocks. As discussed at length in Chapter 12, the probability of
Sams, 1995, pointed out that the minimum detectable a variable, mini-EDZ in the walls of a well that penetrates
layer thickness using standard sonic tools may be no inter-bedded strata of different stiffnesses, may mean
less than 15 cm. In the limit when the layers become accentuated fluctuation of velocities when sonic logging,
thin with respect to the wavelength, the particular layer if the well is of sufficient depth. Both Vp and an associated
will be transparent. Standard logs will tend to under- Qseis, are intuitively likely to move in the direction of
estimate the effects of layering on measurements of attenu- lower values if the mini-EDZ takes the form of shear frac-
ation and anisotropy. Consequently, intrinsic attenuation turing and additional joint and fissure displacements and
and dispersion may be incorrectly assessed. loosening, as it apparently can in shale, in the so-called
Any distorted frequency dependence of the intrinsic ‘alteration’ zone.
attenuation would subsequently adversely affect estimates Both parameters may move in the opposite direction of
of the physical rock properties and of the contained higher values, when the tangential stress concentration
fluids. This is because the intrinsic attenuation is (e.g. 3 H – h) makes a positive contribution to the
380 Rock quality, seismic velocity, attenuation and anisotropy

‘tightness’ of both a stiff matrix (like limestone) and any


local joint structure in these stiffer facies.
The velocity-facies definition given in Figure 14.12,
with implicit and definitive separation of shales, sand-
stones and limestones, leads one to suggest the poten-
tial for facies or strength based estimates of Qseis, using
such differentiated velocities. With appropriate ranges
of uniaxial strengths and velocities for the three rock
types (shale, sandstone, limestone), one might arrive at
‘representative’ Qseis values of 6–7, 10–15, and 20–40 for
the less frequent limestone layers, using the modulus-
model of Figure 13.60.
The key question is then whether the inter-bedded
mix of facies will create a ‘weighted average’ response,
concerning the high frequency, best definition of layer-
ing when well logging. In Chapter 10, where Qseis
was treated in detail, one may note from Figure 10.34,
from the multi-frequency investigations of Sams et al.,
1997 in these same experimental boreholes, that the
8–24 kHz sonic logging gave the lowest Qp range of about
6–14, with a mean Qp of 10.4 and a mean Vp of
3.48 km/s. (See Table 10.4). These fit the modulus-model
predictions.
Figure 14.13, a virtual copy of Figure 10.35, is
repeated here with the important Qp scale firmly in
place, to supplement the somewhat misleading ‘1000/Q’
typically found in numerous publications. If Qseis is
accepted as potentially being ‘a rock mass property’ (i.e.
pores and joints and saturating fluid included), it
apparently helps to give a very believable range of Qp of
about 5 to 12, with sonic velocities ranging from 3 to
4 km/s as shown in the figure.
The lower frequencies of cross-hole of 0.2–2.3 kHz,
gave a mean Qp of 15.7. Interestingly, the lowest fre-
quency 30–280 Hz VSP, giving poorest definition of
the fine inter-layering, gave a mean Qp of 31.3. At first Figure 14.13 Sams et al., 1997 multi-frequency results for four
sight it would appear possible that this is because of the scales of seismic illumination. The highest frequency
dominant effect of higher velocity layers (limestone) sonic logging, giving a form of weighted-average of
and the relative invisibility of the low velocity layers to the numerous rock layers present, seems to be quite
longer wave-lengths, thereby ‘biasing’ the weighted- well matched by the modulus-model of Figure 13.60.
average picture that applies most fully at higher fre-
quency when ‘all’ layers are illuminated. However, in
relation to velocity, this argument does not hold, as 14.4.1 Larger scale differentiation
Figure 14.13 shows VSP velocities to be somewhat of facies
lower (range of approx. 2.8 to 3.6 km/s). The higher
core velocities, ranging from 3.4 to 4.4 km/s, are for Larger scale differentiation of facies described by Kabir
matrix without bedding or cross-bedding joints, natur- and Verschuur, 2000, is shown in Figure 14.14. In this
ally giving higher Qp, they do not fit the modulus- synthetic Picrocol data set, which nevertheless was based
model well since they are not representative of the on real data shot over a salt dome, the model assumed
jointed, bedded rock mass. constant velocity layers, which was a simplification.
P-waves for characterising fractured reservoirs 381

Figure 14.14 Estimated velocity models based on an IFP Picrocol data set from a southern North Sea salt dome and surrounds. Note lateral
velocity variations in individual layers. Kabir and Verschuur, 2000. Reproduced by kind permission.

Nevertheless each layer was described with an average Fracturing or jointing caused by anticlinal structures
velocity and a vertical velocity gradient. Overburden can be a key to effective drainage of oil and gas from a
stress and fluid pressure were assumed to increase lin- porous but low permeability matrix such as tight sand-
early with depth. The authors discussed the sometime stone or chalk. Depending on the efficiency of the cap-
need for lateral velocity gradient in addition to vertical rock, often shales, there may also be over-pressured shale
velocity gradient, in interpreting variations. They sug- layers in the reservoir. Fracturing or jointing causing
gested the use of well velocities to improve depth con- reduced velocities may be seen as a seismic time-sag.
version and geological interpretations. This can be accentuated by the presence of gas rather
382 Rock quality, seismic velocity, attenuation and anisotropy

than oil, and can also be accentuated by the effect of fluid. Subtraction of seismograms for the down-going
over-pressure. and the up-going pulses leaves an enhanced reflection.
(Christie et al., 1995).
There is a growing trend to instrument some petrol-
14.5 Detecting anisotropy caused eum wells on a ‘permanent’ basis, so that 4D seismic can
by fractures with multi-azimuth be used ‘more easily’, to monitor changes bought about
VSP by different water-flood and production practices, using
rock physics principles to assist in the interpretation. As
In the same way that we had to reject earlier assump- we saw in Chapter 10, the in-well installation practice
tions about elastic isotropic rock mechanics behaviour, apparently started in the late 80’s and early 90’s in the
the sedimentary rocks containing hydrocarbons have case of permanent installations in deep holes adjacent to
also proved to be neither isotropic nor homogeneous, the San Andreas fault in California.
but heterogeneous and anisotropic. Besides detecting
azimuthal velocity anisotropy, signalling aligned fractur-
ing and/or anisotropic rock stress, there are now needs 14.5.1 Fracture azimuth and
for spatial resolution of variable structure, and also for stress azimuth from P-wave
resolution of temporal changes. A more recent challenge surveys
is the resolution of azimuthal variation in attenuation as
an indicator of horizontal permeability anisotropy. The original goal of a seismic study reported by Lynn
All seismic data are now known to vary with the off- et al., 1999a, for the US Department of Energy, was to
set (angle in relation to the well), and with the azimuth. evaluate and map fracture azimuth and relative fracture
Since all seismic data also varies with frequency, there is density throughout a naturally-fractured gas reservoir
an increasing recognition that 3D multi-component, interval. P-waves travelling in the plane parallel to aligned
multi-mode and multi-azimuth acquisition may be vertical fractures (thus giving transverse isotropy with a
required and economically justified. The seismic wave- horizontal axis – known as HTI media), have higher
length at which the measurement is made, determines velocity than the P-waves travelling perpendicular to
what seismic attributes can be measured, and whether the fractures. The latter direction corresponds to the
the rock looks homogeneous and isotropic or heteroge- high attenuation direction, due to the lower frequency
neous and anisotropic. of the P-waves, and therefore a lower seismic Q. The
According to Lynn, 2004 when the cause of the advantages of using of both travel time and amplitude,
ordered anisotropy replicates itself across different scale or travel time and frequency, for detecting permeability
lengths, or exhibits fractal tendencies, the anisotropy was emphasised.
measured in modern dipole well logs at kilohertz fre- As the authors pointed out, the detection of relative
quencies, may match the anisotropy trends that are crack density and crack orientations was not the final
detected at another scale, with 10–100 kH reflection goal, rather it was the detection of azimuthal anisotropy
seismic. A given set of vertically aligned fractures will that was attributable to structure that controlled the fluid-
therefore cause anisotropy with low frequency measure- flow properties at reservoir scale. With sufficient fracture
ment, signal distortion with mid-frequency measurement, density, connectivity and permeability anisotropy were
and lead to reflections by the highest frequency waves. likely. A threshold value of anisotropy would probably
The practice of VSP (vertical seismic profiling), walk- be involved. The detection of permeability anisotropy
away VSP, multi-azimuth VSP, and numerous other could be considered as one step beyond the detection of
techniques related to VSP are so fundamental in explor- vertically aligned fractures, and/or the detection of
ation that a diagram for illustrating the technique is unequal horizontal stresses.
probably superfluous. Nevertheless, since a figure is The authors emphasised the importance of know-
worth a thousand words: Figure 14.15 shows the princi- ledge of the orientation of maximum horizontal stress,
ple of e.g. walkaway VSP, with surface shot points shown due to the commonly assumed strong correlation between
along a single azimuth, and downhole receiver positions. directionality of reservoir flow and the local, present-
There is also a technique for performing VSP for hori- day orientation of the maximum horizontal stress. Typ-
zontal well sections, using geophones clamped to the for- ically, oriented four-arm calliper logs show a long axis
mation, and hydrophones suspended in the borehole that is oriented parallel to the minimum horizontal stress
P-waves for characterising fractured reservoirs 383

(a) (b)

(c)

Figure 14.15 a) Simple illustration of the principles of walk-away vertical seismic profiling (VSP). The usual practice will be the use of shot
points along multi-azimuth lines, preferably symmetrically distributed, such as in two perpendicular arrays related to the dom-
inant fracturing direction (e.g. Grimm et al., 1999), or as illustrated by b) with a more complete double-orthogonal coverage
for the case of the hypothetical fractured reservoir (Liu, 2003). A walk-around lay-out at constant offset or radius, for improved
characterization of fractured reservoirs has also been used (Horne, 2003). Use of wide aperture layouts, extending on the sur-
face to at least the target depth is recommended, e.g. Lynn et al., 1999. c) Zero, or limited offset VSP with the direct and
reflected wave-paths explained (Christie et al., 1995).

direction, if there is stress-induced break-out. (However, WNW-oriented fracture strike (based on oriented core
as mentioned earlier in this chapter, and discussed in and impression-packer results). The WNW fracture direc-
more detail in Chapters 15 and 16, there may be joint- tion also agreed with the maximum stress indicated by
or fracture-related reasons for a careful evaluation of this break-out analysis from calliper logs. A plausible explan-
commonly-held viewpoint.) ation of enhanced dilation of the WNW fracture set, due
Interestingly for later discussion of this topic, the to structural flexure of the anticlinal structure, mentioned
authors mention the ‘problem’ of a NW-oriented improved connectivity as a possible explanation of the
production trend at a neighbouring well, yet with a (roughly 20°?) rotated NW production trend.
384 Rock quality, seismic velocity, attenuation and anisotropy

Analyses at the Rulison Field by Lynn et al., 1999b, fracture direction (based on structural geology). This
using just two azimuthal bins of data, also showed a ‘dis- method was used in order to construct azimuthally dif-
crepancy’ of orientations: a similar present day stress as ferentiated, and azimuthally limited seismic attributes,
above (N 70–80° W), but principle P-wave velocity that would highlight the formation’s fracture and
directions of N30W and N60E. The authors continued matrix response, respectively.
to assume that the direction of H max would be the A rough measure of fracture anisotropy was given by
direction of open cracks or fractures, but did not dis- the resulting P-wave interval velocities, measured in two
cuss the possibility of the interaction of e.g. two sets of such azimuthal bins. (Two-azimuthal binning is however
joints or fractures, or the possibility of conjugate shear- blind to anisotropy near 45° to the principal chosen
ing as an explanation of ‘open’ fractures. directions). The authors found that when the anisotropy
This topic is treated in detail in Chapter 16, following exceeded 5 to 10% using this simpler approach, there
analysis of the shear strength of rock joints and fractures, was good agreement with the principle directions obtained
and analysis of what has been learned by Zoback and co- by more sophisticated methods. There was correspond-
workers, about water conducting joint (or minor fault) ingly robust correlation and identification of zones of
directions relative to H max directions. Their findings high fracture density and permeability.
should have special relevance in petroleum reservoirs The authors reported that reflectivity and frequency
where rock strengths are limited, yet fractures are assumed were also anisotropic, and actually had better correlations
to be ‘open’, despite ␴h min magnitudes of tens of MPa. with gas productivity than P-wave velocity anisotropy. Ray
In Chapter 16, (and Barton, 2005), a further reason paths travelling parallel to fractures giving increase in fre-
is discussed for such minor angular discrepancies, namely quency may have been due to squirt flow mechanisms
a dilation-related contrary-rotation of fluid lenses parallel to the fractures. The authors used the neural net-
contra rock-to-rock contacting asperities, when non- work technique (with 85% success), to infer commer-
planar joints (or fractures), are actually under signifi- cial prospectivity over undrilled areas, using the available
cant shear stress, and therefore significantly ‘open’. This geological, geophysical and engineering attributes.
geometric effect could possibly rotate shear wave split- The authors advised full-azimuth 3-D P-wave sur-
ting mechanisms. veys for such analyses, with maximum offsets equal to,
Interesting observations on the subject of ‘open’ fracture or greater than target depth, using azimuthally isotropic
orientations were made by Laubach et al., 2002. They source and receiver arrays. Naturally they also recom-
cited comparisons of measured stress directions and orien- mended processing in as many azimuths as allowed by
tations of open, flow-controlling fractures that showed cost, with independent velocity analyses in each azimuth.
that open fractures in the sub-surface were not necessarily An admirably detailed integration of structural orien-
parallel to maximum compressive stress ( H max). Fractures tation data with near-offset VSP at the Conoco
perpendicular to this direction could also be ‘open’ if par- Borehole Test Facility, was given by Queen and Rizer,
tially filled with mineral cements, and for this reason, 1990, based on the surface outcrop joint orientations,
sealed fractures parallel to H max were numerous. rock fabric orientation (from point-loading), and from
They pointed out that a determining factor for fluid the results of joint orientation using a borehole tele-
flow was the degree of mineral cements deposited within viewer and oriented core. Figure 14.16 shows the prin-
the fractures, either ‘at the time of fracturing’ (synkine- cipal orientation data at these different scales. The two
matic), or as post-kinematic cements precipitated after joint sets seen at the surface proved not to be as consist-
fractures ceased opening. They suggested, from experi- ent as expected with the three sets of jointing in the
ences in both compressional and extensional provinces, sub-surface. The quite consistently oriented point-load
with production data from 2,400 to 6,400 m depth, fracturing traces could not apparently be related to any
that the divergence between H max and ‘open’ fractures observable microcrack directions or to fabric, and
demonstrably contributing to flow was ‘from a few would then seem likely to have some relation to resid-
degrees to 90 degrees’. ual stresses in the samples.
Grimm et al., 1999 also chose a relatively simple The seismic anisotropy measurements consisted of
(two-azimuth bin) method of estimating the spatial dis- near-offset VSP in twenty three 30 m intervals from
tribution of gas-producing natural fractures in a tight 183 to 853 m depth, and azimuthal VSP uniformly
gas reservoir. This relatively large P-wave survey was spaced at 15° intervals around a 120° quadrant, with
divided into two volumes, with ray-paths parallel and nine sources on a 290 m radius centred on the Conoco
perpendicular (/45°) to the assumed dominant 33–1 well.
P-waves for characterising fractured reservoirs 385

Figure 14.16 Top left: surface outcrop and joint rosette. Top right: superimposed tensile-fracture traces from point-load (steel ball) indentation
testing of the limestone. Middle: left and right: cumulative induced-fracture lengths and orientations from 35 tests on oriented sur-
face samples of the limestone, and from 9 tests on oriented indurated shale core from 734 m depth down the Conoco 33–4 well.
Bottom left: Joint rosette concerning accumulative length of sub-surface jointing of given strike, interpreted from BHTV and
oriented core (with bias against non-vertical jointing). Bottom right: Polar histogram for all levels and azimuths of all the nine-
component VSP data, with lengths weighted by travel time. Queen and Rizer, 1990.
386 Rock quality, seismic velocity, attenuation and anisotropy

The results of shear wave splitting and polarization, from spectral ratios. When the two Q seis values agreed,
from this multi-component survey (strictly also belong- they concluded that velocity dispersion resulted solely
ing in Chapter 15), suggested closest correspondence from absorption. They showed logs of Q p down a 1000 m
with the ‘middle’ set of (ENE) fractures, which was close deep borehole that showed an average difference of Q p
to the BHTV-interpreted break-out analysis indication (dispersion formula) minus Q p (spectral ratio method)
of the (perpendicular) major horizontal stress. of 13. Typical ranges of Q p were 10 to 50 (again remind-
Alternatively, the ENE direction could be the theor- ing one of the expected magnitude of deformation
etical and therefore practical resultant of shear wave split- moduli, as expressed in GPa).
ting influences from the other sets, which strike to each They found that VSP (i.e. seismic) Q p were systemat-
side (NE and ESE). The point load fracturing, perhaps ically smaller than those from sonic Q p, therefore sug-
responding to a residual stress (since no microscopically- gesting a bias between VSP and sonic log Vp that could
visible oriented microcracks, nor aligned fabric were not be explained by intrinsic attenuation alone under a
seen), was also in the conforming ENE direction. The constant Q seis assumption. Besides these observations,
authors were of the opinion that their data, although the authors observed that individual values of Q p or Q s
consistent, did not have enough resolution for detailed changed ‘erratically with depth’ unless depth averaging
engineering analysis of fluid flow. was used. The authors suggested that the validity of
seismic Q ‘does not always correlate with the data qual-
ity or with the rocks themselves’.
14.5.2 Sonic log and VSP Since there is a potential relation between rock qual-
dispersion effects and erratic ity Q c and seismic Q p via the empirically derived ‘static’
seismic Q modulus of deformation (Figures 13.59, 13.60), it
must be emphasised that the rock mass quality Q-value
The authors De et al., 1994, from Chevron, compared also usually varies rapidly down any given drill-core,
P- and S-wave velocities and seismic quality factors (Q p unless rock mass conditions are unusually uniform. So
and Q s) using vertical seismic profiling (VSP) and sonic on that basis it would not then be surprising to see
log measurements in five wells, which were situated in ‘erratically’ varying Q p or Q s with depth.
California, Texas and Alberta. The expected bias (VSP An example of rock mass (core-logged) Q-value vari-
transit times were greater than sonic log times) were ations and an alternative rock mass rating RMR from
attributed to normal velocity dispersion, due to which Bieniawski, 1989, down a recovered core is shown in
higher frequency (sonic) waves travel at higher veloci- Figure 14.17. It is unfortunate that such logs of rock
ties than lower frequency seismic waves. Differences in quality are never (?) a part of the geophysicists’ report-
average P-wave travel times ranged from 2.5% to 7% in ing of seismic attributes. There is after all, a strong
the different wells, giving velocity differences between empirical relationship between Q rock and Vp, and there
the two methods that were consistently in the direction is an implicit similarity in the variable values of Q rock
predicted by dispersion. with both rock quality (obviously), and also with depth,
The authors discussed additional potential causes of as we saw in Chapter 10.
the systematic velocity differences, citing local stress
concentrations around the boreholes, altered zones and
velocity anisotropy or lateral inhomogeneity. In fact 14.6 Dispersion as an alternative
stress concentrations around the wells, will cause a mag- method of characterization
nification of (tangential) stress, in the same direction as
major principal stress, and diminution of (tangential) Seismic attenuation has come to be recognised as poten-
stress in the perpendicular direction. If these effects are tially very sensitive to reservoir properties. This is because
strong enough in relation to rock strength, shear sur- of its sensitivity to fractures, joints or bedding planes, and
faces may develop, first giving break-out, subsequently in turn, due to sensitivity to changes of effective stress.
a log-spiral-fractured discontinuum. Various forms of Attenuation levels are also sensitive to the saturating fluid
borehole wall disturbance are possible, as emphasised and petro-physical properties. However, as emphasised
in Chapter 12. by Hackert et al., 2001, field measurements of in situ
De et al., 1994, calculated seismic quality factors (Q p attenuation are complicated due to reflections, geomet-
and Q s) both from a velocity dispersion formula and rical spreading losses and varying formation stiffness.
P-waves for characterising fractured reservoirs 387

California. They found that high dispersion (and low


Q seis) values correlated with thin sand and carbonate
beds within the shale. These beds were at least ten times
as permeable as the host shale formation. A velocity and
density log for the part of the formation (3950 to
4300 ft depth) is shown in Figure 14.18.
The main Antelope Shale (lower half of log) contained
more than 100 thin sand and carbonate layers: the most
visible ones marked with (S) and (C). These contrasted
with the uppermost Brown shale, and Upper Antelope
shale that contained no sand or carbonate layers.
High frequency measurements differ from low fre-
quency measurement due to elastic scattering and
intrinsic attenuation. Based on work by Marion et al.,
1994, and Brown and Seifert, 1997, the authors of this
reservoir study gave the following equation for rational-
ising their two-frequency approach:

Vsonic  Vwell  Vsc  Vi


(10kHz ) (1kHz )
(14.8)

Figure 14.17 Examples of rock mass quality variability in terms of where Vsc  change of velocity due to elastic scattering,
Q-values and RMR, down a 600 m section of the and Vi  change of velocity due to intrinsic attenuation.
UK Nirex Ltd borehole RCF1, at Sellafield. This The former is related to geometric effects including the
was the planned site for a 500–700 m deep shaft, to natural anisotropy (layering) of the medium, while the
develop a Rock Characterisation Facility, for nuclear
latter is caused by inelasticity of the rock matrix and/or
waste disposal feasibility studies. NGI contract
viscous losses of saturating fluids in compliant pores.
report, 1994. The rock was tuff and ignimbrite, of
(obviously) varying quality due to jointing and
They also defined a term (VES) from the above terms:
faulting. These Q-values, modified to Qc using
UCS, correlate with Vp, Emass and depth. VES  Vsonic  Vsc (14.9)

As can be noted in Figure 14.19, VES is significantly


It has recently been discovered that dispersion, or the
higher than Vwell, particularly in the anisotropic sand
well known dependence of seismic velocity on fre-
and carbonate section.
quency, can be used for reservoir characterization, as
Hackert et al., 2001 conducted forward modelling
dispersion is mathematically related to seismic attenu-
of their cross-well set-up in this layered medium, first
ation. Sams et al., 1997 (reviewed earlier in this chapter,
assuming Q seis values roughly proportional to the
and also in Chapter 10), compared velocities and attenu-
dynamic elastic moduli, as follows:
ation from four frequency regimes: vertical seismic
profiling (VSP), cross-well measurements, sonic log- Shale Q seis  34
ging and core measurements. They established dispersion Sandstone Q seis  66
and attenuation profiles as a function of these different Dolomite Q seis  88
frequency measurements. A distinct peak in attenuation
versus frequency was observed at about 20 kHz, related They used the difference between the VES and Vwell
to a crack relaxation mechanism (Aki and Richards, curves to derive the predicted Vi, which yields the
1980) with realistic crack aspect ratios according to intrinsic Q seis, using a method described by Aki and
the crack model of Jones, 1986. Richards, 1980. The results of this analysis indicated Q seis
Hackert et al., 2001, used a similar multiple-frequency values in the thin sand and fractured carbonate beds of
(cross-well: 1 kHz and sonic log: 10 kHz) investigation to between 10 and 50 (appearing to potentially match our
create logs of intrinsic dispersion and attenuation for a modulus-model logic), while the bulk shale had Q seis 50
shale formation in the Buena Vista Hills reservoir in in general, which appears not to be well matched by
388 Rock quality, seismic velocity, attenuation and anisotropy

Figure 14.18 P-wave velocity and density logs in relation to the location of non-shale layers (sand and carbonate) in the Antelope shale.
Hackert et al., 2001.

briefly described in this section. Azimuthally dependent


variation of P-wave amplitude can be related, both the-
oretically, and clearly in practice, to the presence of frac-
turing, and gives reasonable estimates of the orientation
of the fracturing, particularly if only one set is involved, or
if one set is dominant. Two studies related by Pérez et al.,
1999a and 1999b will be briefly summarized, as they
illustrate the method, and they also compare results with
shear-wave related studies, using C-wave (P to S con-
verted waves), and the shear-wave splitting mechanism.
Although the use of shear waves are theoretically
favoured for fracture set detection, using the shear wave
splitting and polarization mechanism to be described in
Chapter 15, there has apparently been a certain reluc-
tance to use shear waves, due to more expensive acquisi-
tion and more expensive processing routines. For these
reasons, the use of P-waves for fracture set detection and
estimation of orientation, has attracted a lot of interest,
Figure 14.19 Cross-well velocity data compared to computed VES. even though P-wave travel times need to be detected in
Hackert et al., 2001 computed Qseis, finding lower
many directions to obtain the necessary information.
Q seis with multiple sand and fractured carbonate lay-
AVO analysis using P-waves, is based on some prin-
ers as shown earlier in Figure 10.69b.
ciples that can best be illustrated by the practical example,
modulus-model logic, suggesting other attenuation mechan- given by the authors. If seismic data acquisition is con-
isms than fracturing and contained fluid effects. ducted parallel to the fracture orientation, the fractures
will have minimal influence on the reflection properties,
regardless of the angle of incidence, or offset. This is
14.7 AVO and AVOA using P-waves because the P-wave particle motion is parallel to the frac-
for fracture detection tures. If the line is instead oriented more perpendicular
to the fractures, at larger angles of incidence than zero,
AVO (amplitude variation with offset) and AVOA the reflection coefficients will be affected strongly. In
(amplitude variation with offset and azimuth) will be fact at the largest angles of incidence, especially when
P-waves for characterising fractured reservoirs 389

Figure 14.20 Survey geometry for Pérez et al., 1999a and 1999b analyses of 3D AVO and converted (C) wave P-S shear wave splitting analy-
ses. Small inward-facing arrows indicate H max direction interpreted from break-out orientation logs, from wells 16, 17, 20
and 23. The (unconventional) rose diagrams, which have been blackened for clarity, indicate the various fracture orientations,
and their density, based on FMS logs at these boreholes.

perpendicular, the P-wave velocity is expected to also be The authors opinion at this time was that there were
affected by the acoustic properties of the fluids filling few studies that related AVO attributes to fracture or
the fractures. So in the presence of anisotropy, the reflec- crack parameters. However they were convinced that
tion amplitude will vary with offset, due to changed P-wave AVO gradients were affected by fracture-
angle of incidence, and also will change with azimuth induced azimuthal anisotropy. Analysis involved calcu-
(AVOA). lating reflection coefficient curves in two azimuths, one
The authors used data acquired from three, intersect- parallel, the other perpendicular to fracture orientation
ing 10 km, three-component seismic lines, with three (implying the need to know the expected result). They
different azimuths. The three lines intersected at one of found that the AVO gradient was larger for the lines
the wells, where there were results of FMS (formation perpendicular to fractures than for the line parallel to
micro-scanner) and caliper logs for estimating both the fractures: this direction showed a somewhat higher
dominant fracture strike and the direction of H max. reflection coefficient.
The latter was oriented NW-SE, and the authors In a companion paper, Pérez et al., 1999b compared
assumed that the fracture set in approximately this their local 2D AVO study of the fracture effects at the
direction (one of several sets), would be the one most intersection of their three azimuthal lines (with tie-in to
likely to be ‘open’, and therefore most detectable. well data at this point), with more comprehensive 3D
In this study the far offset extended to 3600 m, a bit azimuthal AVO analysis, and with analysis using con-
more than the depth of the target zone: a 35 m thick verted (C) P-S waves, using shear wave splitting and
fractured limestone at approximately 3000 m depth. polarization (see Chapter 15).
This had a P-wave velocity of only 3513 m/s, and an The converted P-S waves have the advantage that
S-wave velocity of 1890 m/s, with a density of 2.5 gm/ they can be generated by compressional ( i.e. explosive)
cm3. In other words it would seem to have been sources, yet are expected to contain the same informa-
extremely well fractured or over-pressured. tion as pure S (or SS) waves. These comprehensive stud-
A geophysical term NMO (normal moveout) is rele- ies were finally used to assist reservoir engineers in
vant here, as the so-called NMO stretch decreased fre- exploiting the fractured limestone layer at 3 km depth,
quency at far offsets, which therefore affected the using horizontal wells oriented perpendicular to the
amplitudes. This distortion was most significant at shal- densest (or the assumed more ‘open’?) fracture set.
low depth and large offset. During processing, the authors Figure 14.20 shows the layout of the three survey
observed a higher frequency (and velocity), along the off- lines relative to the wells, with details of some fracture-
set line that had an azimuth parallel to the assumed dom- strike orientations from FMS logs. Figure 14.21 shows
inant (‘open’) fracture set, which agrees with other studies. the results of two of the three fracture direction analyses,
390 Rock quality, seismic velocity, attenuation and anisotropy

The second fracture orientation plot was derived


from the 3D azimuthal AVO analysis. The ‘independ-
ence’ of the plots was guaranteed by the sequential timing
of the surveys. The authors estimated that the azimuth
of the maximum AVO gradient was 56°, based on a
Rüger, 1996 formula for calculating the reflection coef-
ficients in TIH (transversely isotropic media with a
horizontal symmetry axis due to a vertical fracture set).
Consequently, since this is assumed to be perpendicular
to fracture orientation, the 3D AVO based fracture
azimuth estimate was 146°, which was shown to be as
much as 36° from the interpreted H max direction.
When based on the converted P-S shear wave splitting,
the deviation was less.
This brings us to a topic introduced later, when
reviewing water-flood analyses summarized by Heffer,
2002. Heffer’s results from water flood case records,
may suggest to those with rock mechanics background,
the possibility of conjugate-shearing of joint sets that
are intersected by the H max direction. This topic is
treated in more detail in both Chapters 15 and 16, as it
both contradicts the ‘standard industry’ assumption of
‘open fractures parallel to H max’, at the same time
helping to explain real angular discrepancies between
dominant ‘open’ fracture azimuths, and the perpendi-
cular-to-break-out based H max direction. (Refer to
Figure 16.71 for visualization of this shear mechanism,
and the potentially important fluid lense rotation
phenomenon).
The authors Pérez et al., 1999a and 1999b felt that
their analyses of fracture orientation trends from 3D
AVO and from converted P-S (and also from 3D NMO
ellipticity: not shown here due to much ‘poorer’ stress-
direction fit), had ‘captured the regional orientation of
maximum horizontal stress in the field’. They also con-
cluded, significantly, that even though the reservoir had
several fracture systems, most of the applied methods
Figure 14.21 Fracture orientations within the investigation area, picked ‘the one that is closer to the maximum horizontal
projected to the 35 m thick fractured limestone at 3 km stress’. See also Tingay et al., 2005 (Fig. 8c)
depth. Interpretation from a) Converted P-S based As we shall see in Chapter 16, this may not be a
shear wave analysis, b) 3D azimuthal AVO analysis. necessary condition for maximum permeability direc-
Note that the result using converted P-S waves was rel- tion – most likely this will be found sub-parallel to the
atively closer to H max. Pérez et al., 1999b. dominant set of a conjugate pair that has suffered more
shear and dilation than its neighbouring set. At shallow
depth however, the standard ‘industry assumption’
which were described by Pérez et al., 1999b. The upper (also in civil engineering) is more correctly focussed
one is from the converted (C) wave P-S analysis, using on maximum permeability (and Vp) being parallel to
the shear wave splitting and polarization method (qS1 the H max direction. This is because this joint direc-
parallel, and qS2 perpendicular to jointing – see tion has verifiable apertures and permeability near the
Chapter 15). surface.
P-waves for characterising fractured reservoirs 391

(OBS) survey was performed at the Valhall field in


1998. (See details in Figure 14.22, from Hall and
Kendal, 2003). This, and previous 2D-4C studies had
shown the ability of shear waves, derived from converted
P-S waves, to penetrate through a gas cloud and illumin-
ate the target fractured zones, otherwise invisible to
P-waves in these locations.
As part of the seismic studies using the new 3D OBS
data at Valhall, Hall and Kendall, 2000 and subsequently
Hall and Kendall, 2003, addressed the problem of model-
dependence in AVOA analyses of the multi-azimuth
OBS-acquired, vertical P-wave part of the data (i.e. com-
ponent Z in Figure 14.22c).
The authors suggested that the fracture strike direc-
tion could be ambiguous, since the azimuthal variation
in the near-offset AVO gradient, could be positive or
negative, relative to the fracture direction. They found,
unfortunately, that the direction of the most positive
AVO gradient could correspond to either the fracture-
normal, or the fracture strike direction, depending on the
character of the fracturing, i.e. depending on whether
brine-filled or gas-filled – which were the ‘end-member’
models.
They therefore recommended forward modelling in
order to constrain the interpretation of AVOA. The
objective was to predict the near-offset AVO gradient
anisotropy for equivalent medium models, based on the
different elastic properties of the different fracture mod-
els, and compare this with observed AVOA in the data.
The effective medium models were based on the addi-
Figure 14.22 a) Exaggerated vertical cross-section of faulted anticline tional fracture compliance terms (ZN and ZT) allowing
defining the Valhall reservoir. The small-scale fracture poroelastic modelling of the (micro) hydraulic connec-
sets are thought to be related to the domal faulting, tivity between fractures and pore space (as discussed in
which has principal directions of NW-SE and NE-SW. more detail in Chapters 10 and 15).
b) Schematic of orthogonal 3D-4C ocean bottom Log-based data for the fractured chalk were as follows:
cable (OBC) data acquisition over Valhall. Source lines Vp  3134 m/s, Vs  1534 m/s, density 2300 kg/m3.
were 7.8 km long at 600 m spacing, with shots every Permeabilities related to production were an order of
25 m over pairs of receiver lines. Each receiver cable of magnitude higher than core data, due to the significant
6 km length, also at 600 m spacing, had receiver arrays
amount of fracturing. Hall and Kendall, 2003 men-
at 25 m spacing, each with seven four-component sen-
tioned that azimuthal anisotropy was also observed from
sors at 1.5 m spacing in each group. Note: patch shoot-
ing, over pairs of receiver lines, using air guns at 7.5 m
the dipole sonic logging in the producing horizon, using
depth. c) Details of a single, multi-component receiver the flexural shear-wave splitting principle, as described
unit, which are uniformly spaced along the cables, in Chapter 12.
Hall and Kendall, 2003. Their comparison of model and field data suggested
that a large crack density (about 0.1) and thin fractures,
14.7.1 Model dependence of AVOA were needed to match the high anisotropy, and this
fracture orientation matching also suggested liquid-filled rather than gas-
filled fractures, unless aspect ratios were 0.00025.
As described later in this chapter, a 3D-4C (three- Low matrix permeability was also suggested. They
dimensional, four-component) ocean bottom seismic considered a most important finding was that the high
392 Rock quality, seismic velocity, attenuation and anisotropy

Figure 14.23 Fracture pattern map, assuming that the high levels of observed anisotropy could only be modelled if the most positive near-
offset AVO gradient was in the direction perpendicular to dominant fracturing. Hall and Kendal, 2003.

levels of observed anisotropy could only be modelled if the mechanism being modelled, i.e. squirt flow attenu-
the most positive near-offset AVO gradient was in the ation would actually be governed by a different aspect
direction perpendicular to dominant fracturing. Hall ratio to that determining the fracture compliances – if
and Kendal 2003 emphasised that in other situations, these mechanisms were being modelled in poroelastic
the reverse could be true, i.e. the direction of the most models that required aspect ratio as input.
negative AVO gradient could be in the direction per- The standard assumption that maximum permeabil-
pendicular to the dominant fracturing. ity in a rock mass tends to be parallel to a dominant set
The principal fracturing directions interpreted from of vertical joints or fractures, which themselves usually
the AVOA studies described, were NNW-SSE to N-S, trend parallel to the maximum horizontal stress, is a
the former nearly resembling the ‘compartmental- simply understood concept for which there is also the-
like’ NW-SE component of faulting, as shown in oretical support (e.g. Sayers, 1990). The arguments
Figure 14.23. clearly extend also to the P-wave and S-wave velocities.
The concept and the theory are defensible, when the
rock is of sufficient strength in relation to the effective
14.7.2 Conjugate joint or fracture normal stress, to provide (partly) ‘open’ joint apertures
sets also cause anisotropy in the major stress-parallel direction. This generally
applies at least to near-surface rocks, and it may apply
From experiences of domal-structure jointing at Ekofisk, at reservoir depths in the case of harder rock types.
one may pose a tentative question: could the above However, if stress-closure modelling that is based
images of dominant, fracturing directions actually be on empirical rock mechanics data, can demonstrate
images of the similar strike of conjugate, steeply dipping virtual closure of the set under discussion, at effective
sets? If so, then an important mechanism of compaction normal stress levels of several tens of MPa, then other
and production-maintenance could also be at work, mechanisms are likely to be operating, in order to
namely down-dip shearing, despite the one-dimensional explain a viable production rate, assuming that matrix
strain boundary condition. (Barton et al., 1988). permeability is insufficient to explain production, as for
The modified apertures caused by slight shear- instance at both the Valhall and Ekofisk jointed chalk
induced dilation, could perhaps have influence on the reservoirs.
above aspect ratio assumptions, where the inequality of The conjugate shear mechanism, and an important
physical aperture (E) and conducting aperture (e) detail of this mechanism for the common case of non-
(Barton et al., 1985), actually results in two possible planar joint or fracture surfaces, is introduced in
aspect ratios, whose individual relevance may depend on Chapter 15, and further quantified in Chapter 16, with
P-waves for characterising fractured reservoirs 393

Figure 14.24 a) Two synthetic models of jointed or fractured reservoirs, with one set, or two intersecting sets of joints. The modelling of
overlying shale (Vp  3700 m/s) and reservoir matrix (Vp  5877 m/s and density 2.4 gm/cm3), was supplemented by com-
pliance additions to account for the joints, but with neglect of complex joint-set interaction (as modelled in UDEC).
b) Azimuthal P-wave reflection coefficients as a function of angle of wave incidence at four selected azimuths for models
1 and 2. c) Equivalent coefficients for split shear waves: P-SV for model 1 and P-SH for model 2. Note greater azimuthal sen-
sitivity of a single set of joints, and greater azimuthal separation of the reflection coefficients for the case of polarized shear
waves. Sensitivity of P-waves to azimuth is limited, below angles of incidence of 25°. Chen et al., 2005.

particular reference to the ‘shearing critical crust’ con- Some of the theoretical assumptions concerning the
cept of Zoback and co-workers. So far little recognition ratio of compliances being 1.0 for the case of ‘open, gas-
of the importance of this mechanism is evident in reser- filled’ fractures seem not to match engineering concepts,
voir geophysics literature. due to the entirely different mechanisms involved in clos-
Sayers and Dean, 2001, and also Chen et al., 2005, ure and shear. However a lower ratio of ZN/ZT (or Bn/Bt)
addressed the question of the effect of additional fracture for the case of clay filled fractures seems reasonable.
sets, in particular non-orthogonal vertical sets, on the Chen et al., 2005 presented reflection coefficients as
AVO response. The first authors showed sinusoidal- a function of both angle of incidence and azimuth for
type P-wave reflection coefficient trends, when plotted two very ‘tangible’ images of single-joint set and two-
versus azimuth (0 to 360°), with variations depending joint set models, as illustrated in Figure 14.24. The
on the assumptions for the ratios of fracture normal authors set up synthetic jointed reservoir models with
and shear compliance, and whether these ratios were the model 1 and model 2 jointing, intending to have a
equal or unequal for the different sets. fracture density of 0.1 in each case.
394 Rock quality, seismic velocity, attenuation and anisotropy

14.7.3 VP anisotropy caused by going from a vapour-saturated (low pore pressure) condi-
faulting tion to a liquid-saturated (high pore pressure) condition.
Water-dominated geothermal fields tend to have higher
Concerning anisotropy in the neighbourhood of faults, values of  (i.e. 0.2–0.3), while steam-dominated fields
Williams and Jenner, 2002, warned of the possibility for have lower values of  (i.e. 0.15–0.2). NW-SE and NE-
multiple fracture directions, with the fast velocity no SW trending structures associated with reservoir fluid
longer equal to the matrix or bulk rock velocity. The nor- channelling were suggested by the anisotropic distribu-
mal elliptical Vfast and Vslow distribution is then replaced tion of , according to Simiyu, 2000.
by superimposed multiple ellipses, which have the effect
of reducing the observed velocity, and the previous direc-
tionality with a single set of joints or fractures will be lost. 14.8 4C four-component acquisition
Due also to rapid changes in fracture frequency, rapid of seismic including C-waves
changes in velocity are also seen – in fact responding to
the rapid changes in rock mass quality Q close to, and There is a multitude of technical jargon in the geo-
across faults, as frequently observed in tunnelling, and physical industry, not least because of the advanced
when logging fault-zone core for rock quality purposes. developments, the large number of practitioners, and the
The authors showed a velocity difference tomogram complex processes of seismic data gathering, and meth-
of Vfast – Vslow, with velocity differences of up to ods of interpretation. One of the simpler sources of
150 m/s in the neighbourhood of known faulting, due confusion is double use of C – which singly used refers
not to bulk velocity changes, but due to changes in frac- to converted waves (i.e. P-wave converting to S-wave
turing frequency and directions. Since even 2-D offset mode, at the sea floor), and the 4C term meaning four-
lines cross the earth’s azimuthal velocity field, a case was component seismic recordings, typically using ocean-
made for dense sampling near fault and fracture zones. bottom-cable (OBC).
Mostly, azimuthally varying P-wave velocities are attrib- A helpful diagram, illustrating each term, and also pre-
uted to fractures, but the authors suggested that stress- senting an important ‘new’ seismic acquisition method,
aligned microcracks, as always suggested by Crampin, was found in Yuan, 2001, who analysed four-component
may also be responsible, and that the fractures did not seafloor data to determine the possible presence of verti-
need to be ‘open’. cal (TIV) or horizontal (TIH) transverse isotropy, in the
presence of mode-converted shear waves.
The illustrated acquisition technique consists in prin-
14.7.4 Poisson’s ratio anisotropy ciple of implanting four-component (4C) sensors into
caused by fracturing the seabed. These consist of one hydrophone, one verti-
cal geophone, one in-line horizontal geophone, and one
A useful contribution to the question of seismic cross-line horizontal geophone. (See also Figure 14.22c).
anisotropy caused by fracturing is found in a descrip- Conventional air gun arrays (P-wave generating sources)
tion of geothermal development in the Olkaria field of are used, towed by a shooting vessel, while a recording
the Kenya rift. This incipient caldera structure has had vessel stays above the receiver array. The receiver (OBC)
a number of injection episodes. It was monitored with array must of course be relocated, to give multiple-
a seismic network comprising 18 stations. Simiyu, 2000 azimuth data if not already installed in multiple azimuth
gave Vp, Vp/Vs, and  data from 2613 well-located directions. There is a potential tendency for 3D OBC to
micro-earthquakes, 45 quarry blasts and 25 calibration give ‘patchy coverage’ in offset and azimuth (e.g. Hall
shots. Velocity-depth data, extending beyond 4 km and Kendall, 2003), though not in the case of the exten-
depth, for six geothermal fields in the area (identified sive survey illustrated in Figure 14.22.
by initials) are reproduced in Figure 14.25. The well An interesting proposal by Thomsen, 2002, based on
distributed stations made possible the presentation of imminent plans at BP, was to ‘fuse’ 4D (repeated surveys
less frequently seen, azimuthally dependent dynamic for reservoir management), with 4C. Although 4C tech-
Poisson’s ratio data, as shown in Figure 14.26. nology had proved an economic success, e.g. for imaging
For the case of geothermal reservoirs, Vp/Vs and  beneath gas clouds, and helping to delineate anisotropic
are more controlled by Vp than Vs due to the various structure, it had not according to Thomsen caused a revo-
fluid-phases in the reservoir. Vp/Vs ratios increase when lution. He also considered that 4D, though very useful
P-waves for characterising fractured reservoirs 395

Figure 14.25 Vp and Vs-depth trends and Vp/Vs histograms for six geothermal fields in the Kenya Rift. Simiyu, 2000.
396 Rock quality, seismic velocity, attenuation and anisotropy

for enhancing reservoir management, was expensive as done frequently. The key was to accept the investment
each new survey cost as much as the previous one. of permanently installed receivers in the seafloor, so
Thomsen’s vision was to make subsequent 4D surveys that the marginal cost increase for making 4D receivers
much cheaper than the first, so that re-shoots could be also 4C, would be born by the economic gains made
from ‘dozens’ of 4D re-shoots. Management needed to
become confident in the gains to be made from fre-
quent 4D4C surveys.
Such a philosophy, based on huge investment until
prices were driven down by marketplace economics,
could revolutionize the practice of reservoir manage-
ment according to Thomsen, with ‘history matching’
replaced by ‘parameter estimation’ and prediction of
performance. The consequence of frequent full-field
4D4C re-shoots, providing full-field estimates of all
required reservoir parameters, would result in much
more efficient exploitation of reserves, a production
increment coming sooner, and also possibly larger, than
discovery and exploitation of new fields.
Regarding penetration of gas clouds, referred to by
Thomsen, Granger et al., 2000 referred to the fact that
even 2D-4C surveys at the Valhall reservoir in the
North Sea, had been able to penetrate through a gas
cloud using the converted or C-waves, since the resultant
shear-wave component could illuminate a target other-
wise invisible to P-waves. In 1998, a 3D-4C seismic
survey was performed at Valhall with the intention of
improving the structural imaging of the crest of the field,
and for establishing the potential for jointly using P-wave
Figure 14.26 Dynamic Poisson’s ratio as a function of azimuth and C-wave data for reservoir characterization, lithology
and depth. Simiyu, 2000. prediction, and for stress and fracture orientation.

Figure 14.27 Configuration of four-component (4C) seafloor seismic, for P-wave and mode-converted S-wave acquisition, using ocean-bot-
tom-cables (OBC). Companie Général de Géophysique (CGG) diagram, from Yuan, 2001.
P-waves for characterising fractured reservoirs 397

The P-wave part of this work was briefly reviewed ear- through finite compliances, to attenuation, as already
lier, see Figures 14.22 and 14.23. seen in Chapter 10. This emphasises the need for both
velocity and attenuation monitoring in 4D seismic.
In addressing stress sensitivity questions related to
14.9 4D seismic monitoring of 4D seismic, MacBeth, 2004, developed semi-empirical
reservoirs pressure-sensitive bulk-moduli and shear-moduli for-
mulations using the concept of excess compliance, which
In recent years, many more oil companies have been was designed to capture all categories of weakness in the
utilising Ocean Bottom Cable (OBC) acquisition meth- rock, specifically in numerous types of sandstones. He
ods to make repeated three-dimensional measurements found that the bulk and shear modulus pressure-sensi-
over time, so-called four-dimensional or 4D seismic. (e.g. tivities (i.e. sensitivities to effective stress change), lay
Bull-Gjertsen, 1998). This has made it possible to track between 1 and 10% per MPa.
reservoir depletion phenomena, such as changing pore This meant that a 5 MPa pore pressure decline in a
pressure, and especially to track water-injection fronts in reservoir could produce from 5 to 50% increase in bulk
water flood treatments. Experiences from several North modulus, depending on sand/sandstone type. Rocks of
Sea oil and gas reservoirs have shown that the higher higher porosity tended to show higher pressure sensitiv-
velocity of reservoir zones saturated with water allows ities, and the reverse behaviour was also shown, as intui-
operators to register where the oil-water contact is mov- tively expected. MacBeth pointed out several potential
ing, and where oil may be by-passed by the water flooding. problem areas related with using ultrasonic core meas-
In this section we will see a brief selection of uses for urements for interpreting changes of reservoir seismic
4D seismic, including the mapping of oil saturation velocities in 4D seismic monitoring. Some of these are
changes over time, the monitoring of water-flooding, summarized below:
and the detection and quantification of reservoir com-
paction, and even near-surface, sub-sea subsidence. 1. In saturated rock, higher velocities are generally
Reservoir monitoring with 4D seismic in its most basic measured at higher frequencies.
form is the inversion of seismic data to obtain (dynamic) 2. Saturated samples containing microcracks would
reservoir properties, which can subsequently be used to project a lower stress sensitivity with laboratory
predict pore pressure change at a distance from the wells ultrasonics, than with seismic waves.
due to the assumed, laboratory-sample-based effective 3. Cores loaded back to their original in situ stress
stress sensitivity of the reservoir rocks. While there is high state do not recover their original velocities, and
sensitivity to effective stress in shallow reservoirs, the typ- due also to the possible microcracking caused by
ical ‘plateau’ that may be reached at high effective stresses sampling, may show increased sensitivity to effect-
(i.e. beyond roughly 25 to 50 MPa, depending on rock ive stress relative to in situ.
type), suggests that the fluid compressibility effects may 4. On the other hand, jointing is not a part of the labora-
become more important at greater depth. tory sample response.
5. Cores may be selected from the most productive
and competent part of the reservoir.
14.9.1 Possible limitations of some 6. Subsequent core-to-log correlations, and interpret-
rock physics data ation using long wave-length seismic averages may
cause underestimation of pressure sensitivity.
There is unfortunately a basic complication that the
Most of the rocks studied by MacBeth were predicted
more compressible grain-boundary cracks with their
to have only 1% per MPa sensitivity to stress change (at
low aspect ratios may be partly the result of stress
24 MPa effective stress). However, elevated temperature
unloading when drilled and bought to the surface (Holt
testing may be needed; see Chapter 16.
et al, 1996, 1997, Nes et al., 2000), and it is such sam-
ples that are the basis for much of the collective assump-
tion of a given stress sensitivity. 14.9.2 Oil saturation mapping with
There is in addition the possibility that joint sets that 4D seismic
are under lower levels of effective stress may still be
contributing to some of the (anisotropic) sensitivity to An impressive and very clear indication of the utility of
effective stress change, and they certainly contribute, 4D time-lapse seismic analysis of producing reservoirs
398 Rock quality, seismic velocity, attenuation and anisotropy

was provided by Eide et al., 2002. The authors described


quantitative mapping of changes in oil saturation over a
30 km2 area in the northern part of the Gullfaks field in
the North Sea. A base survey dating from 1985 before
the start of production, was supplemented by monitoring
surveys in 1996 and 1999, for assisting well planning,
and these reportedly had a large economic impact on
production.
The oil saturation mapping project utilised a rock
physics model to compute changing density from P-wave
and S-wave velocities at the three survey times, with
porosity, percentage of clay, saturation and pressure effects
extracted from a flow simulator. Synthetic seismic ampli-
tude and amplitude difference volumes were generated by
forward modelling, based on time-dependent elastic
parameters. Figure 14.28 Uncertainty in estimated velocity and compaction
magnitudes, as a function of reservoir depths, with
The change in reflection strength at the top of the
assumed excellent 4D data yielding a time-shift error
reservoir horizon, and the change in oil saturation in
estimate of only 0.5 ms (solid-lines). With poorer
the upper cells of the reservoir simulation model, using data, and 1.5 ms error (dotted-lines). Maximum off-
segments with high correlation of seismic changes and set assumed limited to 3000 m. Compaction: black,
saturation changes, were used to generate maps of sat- Velocity: grey. Landrø and Stammeijer, 2002.
uration change. Statistical analysis indicated that the
relationship between the change in reflection strength estimated 6 m of additional compaction at 3 km depth.
and the oil saturation change, depended on the height of The authors related time shift, thickness change, and
the original oil column. velocity change with the following normalized relation:
Coloured maps of saturation change from 1985 to
1996, were contrasted with those from 1996 to 1999, t z v
  (14.10)
showing as expected, significantly larger areas and t z v
stronger changes (i.e. So approximately as high as 0.35
to 0.5 in 4 to 5 regions) in the first 11-year period, The authors assumed a 4D time-shift error of 0.5 ms
compared to the less well distributed, small areas and at near-offset, and 2 ms at far-offset, and a reservoir at
small changes in the second, 4-year period. Probability 3000 m depth, with 9 m of compaction. A major chal-
maps of large, medium and low saturation changes lenge here, was to discriminate between compaction
based on a stochastic simulation were also developed. and velocity changes, both in the 300 m thick reservoir,
and in the 3 km of overburden. Figure 14.28 shows a
set of their compaction and velocity-change estimates
14.10 4D monitoring of compaction for a more general case. As reservoir depth increased,
and porosity at Ekofisk the magnitude of uncertainty increased for both
components.
Landrø and Stammeijer, 2002, described two seismic Smith et al., 2002 described some of the detailed
methods for monitoring compacting reservoirs, based on operating problems caused by the Ekofisk subsidence,
seismic data acquired in 4D surveys at Ekofisk. One was including the difficulty of extrapolating seismic depth
based on prestack travel time changes, the other on post- conversion from known horizon-depths at existing ver-
stack travel time and amplitude changes. Velocity-porosity tical wells, to predict depth structure away from the
relationships were not required. The authors emphasised wells. Two hundred wells in the main field area, giving
that 4D time-shifts in a compacting reservoir were a com- excellent depth control, contrasted with limited down-
bined result of increased velocity due to compaction, and flank control, above this large domal structure. The
reduced layer thickness, i.e. reinforcing effects. They authors found that simply scaling seismic interval
referred to measurements of time-shifts of as much as velocities to well interval velocities, suggested larger
12–16 ms, between 1989 and 1999, related to an structural variation than expected. Extrapolated depths
P-waves for characterising fractured reservoirs 399

Figure 14.29 Comparison of a) detailed seismic-based (4D time-shift) compaction interpretation (with adjustment for the velocity reduc-
tion caused by a subsiding overburden), with b) geomechanics-based one-dimensional strain compaction model, that included
porosity reduction due to weakening effect of water saturation, seen in the next figure. Smith et al., 2002. Note gas cloud effect
in centre of seismic model. (See Plate 4). Reproduced by permission from NPF.

down-flank were deeper than nearby exploration wells. paper by Guilbot et al., 2002. The water injection pro-
The authors therefore developed a method for improv- gramme, designed to compensate for the big pressure
ing the seismic depth structure match to the horizontal draw-down following about 20 years of production,
sections of wells that penetrated far down-flank. had side effects, well known from early rock mechanics
The authors mentioned that a straightforward time- testing, of reducing the strength of the chalk.
lapse comparison between the 1989 3D seismic survey, Figure 14.30 shows how this water-weakening was
with the newer 3D seismic survey in 1999, had given tied to assumed porosity reduction, due to accelerated
unrealistically large values for compaction and pore-collapse, based on tests of ‘water-flooded’ triaxi-
subsidence. Subsequently a geomechanics-based model ally confined chalk. In fact the water weakening effect,
was developed, based on vertical uniaxial strain although not positive in terms of arresting settlement of
behaviour for (intact) chalk, that related porosity the sea-bed, has given a strong boost to the compaction
changes to layer thickness changes. The authors men- drive mechanism, resulting in exceptional recovery of
tioned use of a linear porosity-velocity relationship for the reserves.
the chalk, to compute depth conversion from velocity The extensive casing damage to numerous wells at
change. An example of the good general fit between Ekofisk is one set of evidence of discontinuous behav-
the seismic time-shift compaction model, and the iour, due to stretching of the overburden and differen-
geomechanics-based compaction model, is reproduced tial bedding plane slip. A new source of evidence for
in Figure 14.29. (see also Plate 4). discontinuous behaviour during the compaction at
The water-weakening porosity-reduction model used Ekofisk can be seen in the results of the 4D seismic.
at Ekofisk is shown in Figure 14.30, from a companion A 1989 to 1999 ‘time lapse’ comparison (Geo, 2001),
400 Rock quality, seismic velocity, attenuation and anisotropy

the ‘stretching’ of the overburden in response to the 10


years of incremental compaction. Apparently the above-
mentioned, initial over-estimates of compaction given
by the initial 4D comparison of 3D-1989 and 3D-1999
data, did not account for this reduction in velocity.
Large-scale, 2D axi-symmetric modelling with 3 
10 km ‘layered-jointed-and-faulted’ models with UDEC-
MC (Barton et al., 1986, 1988), add support to possible
reasons for a velocity reduction mechanism, as they
showed a lot of locations in the over-burden with hori-
zontal shearing, due to velocity-based moduli-contrasts
between different layers. These events could presum-
ably affect velocities determined from VSP, at most off-
sets. The combination of bedding plane shear and the
stretching caused by the overburden possibly subsiding
nearly as much as the compaction, were presumably also
the inherent source of numerous casing collapses men-
tioned by Ekofisk authors in the past. At the Wilmington
field, where hundreds of casings collapsed, seismic-
(a)
magnitude events were recorded due to bed slip, one of
which measured 24 cm.
Stretching causing slight opening of numerous sub-
vertical, bedding limited joints, and a general horizontal
stress reduction on sub-vertical faults, would presum-
ably only affect velocities determined from wide-offset
(large aperture) VSP.
At Ekofisk there was inevitably a mismatch of 3D
isotropic continuum based subsidence modelling (mostly
giving S/C ratios of about 0.6, and 2D discontinuum mod-
elling, which suggested S/C ratios as high as 0.75, increas-
ing to beyond 0.85 as compaction progressed. (Barton
(b)
et al., 1988). The ‘simpler’ 2D discontinuum modelling
Figure 14.30 The water-weakening model for estimating porosity matched the steep 1985 subsidence bowl better than the
reduction as a function both of vertical effective 3D continuum models, which is understandable con-
stress, and degree of saturation. Guilbot et al., 2002. sidering the 150 km3 of (obviously discontinuous) rock
involved.
appears to show fault related discontinuities in the esti- This total volume of deforming rock, has required
mated ‘tomogram’ of compaction magnitudes, and also billions of dollars of extra investment, due to the late
a somewhat larger compaction (8 m in the central 1980’s need to jack all platforms and risers 6 m, protect
1.5  2.0 km), as compared to the ‘smoother’ continuum the central storage tank with a 100 m diameter concrete
calculations of the geomechanical model reported by wall, and finally establish new platform facilities away
Smith et al., 2002. from the central subsidence during the 1990’s. However,
the compaction is fortuitously giving much more in
return, with greatly increased recovery.
14.10.1 Seismic detection of The huge volume of rock that is deforming, happens
subsidence in the to be some 1015 times larger than the core-plugs used
overburden for uniaxial strain testing. Possibly the high stress levels
are responsible for a relative lack of scale effect, when
It is interesting to note from Smith et al., 2002 that these small tests are used for compaction modelling.
subtle overburden velocity changes had occurred due to Alternatively, the effects of joint deformation in the
P-waves for characterising fractured reservoirs 401

compaction, are in some way ‘cancelling’ a possible


scale effect.

14.10.2 The periodically neglected


joint behaviour at Ekofisk

Ekofisk is a classic jointed (or fractured) reservoir, hav-


ing too low matrix permeability for production without
the jointing. Yet jointing tends to be ‘ignored’ when
modelling, due to the large scales involved (possibly as
much as 15 km3 of chalk, with 300 m thickness). A
more complex aspect that is inevitably missing from
large-scale geomechanical models, which due to size
constraints, are usually isotropic elasto-plastic continuum
in nature, is the actual jointed nature and possible
coupled behaviour of the Ekofisk joints, and their special
behaviour during compaction, especially when stimu-
lated by HTM effects during water-flooding (i.e. cool-
ing, reduced shear strength, greater permeability).
The joints are typically in the form of conjugate sets
of steeply dipping joints, making small block sizes of
e.g. 15 to 40 cm, whose roughly common strike report-
edly ‘rotates’ around the pear-shaped flanks of the anti-
clinal structure. The rotation is associated with the
rotating directions of maximum horizontal stress, also
shown by Teufel and Farrall, 1992, which is reproduced Figure 14.31 The anticline-related rotation of the maximum hori-
in Figure 14.31. zontal stress directions at reservoir levels at Ekofisk.
Two-dimensional distinct element (jointed) model- Teufel and Farrell, 1992 and Bruno and Winterstein,
ling with UDEC-BB performed for the Norwegian Oil 1994. These stress rotations relate also to the rotating
strike of the steeply-dipping conjugate joint direc-
Directory, soon after the seabed subsidence was dis-
tions, which give a rotating trend for principal per-
covered (e.g. Barton et al., 1986, 1988), had shown
meability directions. A rotating trend for P-wave
that down-dip shearing could occur with various joint anisotropy, and for polarized shear waves would also
geometries and with the higher porosities. The one- be expected. (See Chapter 15 for rotation effects
dimensional compaction modelling was performed using measured in the overburden).
the effect of a 20 MPa reduction in pore pressure
experienced prior to water flooding, causing a 20 MPa
increase in effective stress. In subsequent more compre- Barton-Bandis joint strength and stiffness parameters
hensive UDEC-BB modelling at NGI, performed by derived from JRC, JCS and r, were scaled from assumed
Gutierrez in the 1990’s, 24 MPa was used, confirming in situ block sizes, for each set of conjugate joints.
the joint-shearing trends of the earlier modelling. The The less than obvious discontinuum component of the
horizontal stress acting in all these models was, in compaction mechanism, was due to matrix contraction
effect, the rotating major horizontal stress shown in providing ‘space’ for the down-dip shear, despite the one-
Figure 14.31. dimensional (‘roller-boundaries’) constraint. Significantly,
Input data for the 1985 models were obtained from the down-dip shear mechanism caused the development
tests on jointed core recovered from Ekofisk, using joint of a higher ko (ratio of horizontal to vertical stress), than
index testing (e.g. joint wall roughness and tilt tests), in the case for ‘1-D’ compaction of unjointed chalk,
direct shear tests, and coupled-shear-flow-temperature thereby helping to stabilize the mechanism. (Some of this
biaxial tests (CSFT, Makurat et al., 1990), using hot modelling is shown in the joint behaviour related geome-
Ekofisk oil or carbonate-equilibrated sea water. The chanics material, treated in Chapter 16).
402 Rock quality, seismic velocity, attenuation and anisotropy

It seems highly likely that the multitude of (inevitably)


shearing joints in this pear-shaped 9  14 km and 300 m
thick reservoir, may be an important reason for the rapid
weakening by water during water flooding, due to the
huge increases in surface area exposed to water, caused
by the multiple joint surfaces, with strength loss and
increased shear, following mechanisms interpreted at
larger scale by Heffer 2002, reviewed shortly.
The ‘several-micron’ sized apertures modelled in the
1980’s, were apparently almost maintained by shearing
due to joint-surface non-planarity, even without subse-
quent 1990’s and present cooling effects from water-
flooding. Sufficiently permeable jointing (and fractured
rock near faults), will also be potential causes of faster
water break-through to producer wells. Faster break-
through of the water front close to faults was referred to
Figure 14.32 The conventional and frequently occurring fracture
by Guilbot et al., 2002.
and fault opening parallel to the Sh min direction with
These authors reported that about 9 ms of time-shift
water flooding. Heffer et al., 1997, Heffer, 2002.
had occurred at the top of the reservoir in the 10-year
period between the two 3D surveys, but mentioned a
range of 0–20 ms in the water-flood areas. Strongest causing potentially increased velocity, assuming break-
compaction (i.e. largest time-shift) was usually inter- down of the rock frame is not itself a source of velocity
preted where reservoir thickness and porosity was highest. reduction: probably it is initially.
As a possible corollary to this, the above joint-shearing
mechanism was also found to occur with most strength
where modelled porosity was high (e.g. 40%) rather 14.11 Water flood causes joint
than low (e.g. 25%). opening and potential
There are undoubtedly multiple reasons for the shearing
higher compactions where porosity and thickness was
greatest, and also multiple reasons for the continued In water-flooding, for stimulating petroleum production,
reservoir compaction, despite (or because of?) extensive indeed for driving production, there is both a local
water-flooding from 1990. Barton, 2002b suggested increase in pore pressure at the injector wells, and a reduc-
the following: tion in temperature, causing some contraction of the
matrix, both of which help to dilate and indeed create,
● accelerated pore collapse fractures. Gutierrez and Makurat, 1997, have demon-
● joint shearing compaction strated these effects in fully-coupled MHT modelling.
● joint permeability maintenance (due to dilation Heffer, 2002, referred to earlier studies of water-
counteracting effective normal stress increases) flood effects in eighty reservoirs, both fractured and
● water-weakening (accelerated through a vast area of unfractured (Heffer et al., 1993, 1995), and discussed
joint planes) the important question of flood directionality in relation
to the major horizontal stress level. The conventional
Maury et al., 1996 have pointed out the inevitability and expected mechanism of fracture or fault opening
of compaction increase when water-flooding water- exactly in the direction of Sh min, as shown in Figure
sensitive reservoir rocks. Strength reduction actually 14.32, was not in fact so common as one might expect.
applies to the majority of rocks, according to the exper- Figure 14.33 shows the important statistical spread
imental data reviewed by Barton, 1973, but perhaps is of observations, from 33 cases of reservoirs known to be
more extreme for a very porous weak material like chalk. fractured, and from 47 cases thought to be unfractured,
From a change-of-seismic-velocity point of view, the at least prior to the water-flooding.
effect of each of the above mechanisms must be There are in fact subtle and important indications in
extremely complicated, possibly only pore collapse these two ‘joint rosettes’ (in this case fault or fracture
P-waves for characterising fractured reservoirs 403

Unfractured reservoirs Fractured reservoirs


(47 cases) (33 cases)

0 0
360 10 20 360 10 20
350 350
340 30 340 30
320 40 320 40
310 50 310 50
300 60 300 60
290 70 290 70
280 80 280 80
270 90 270 90
260 100 260 100
250 110 250 110
240 120 240 120
230 130 230 130
220 140 220 140
212 150 212 150
200190 170160 200
190 170160
180 Shmax 180

Figure 14.33 Statistical data for water flood directionality effects in relation to Sh max directions in 47 unfractured reservoirs, and in 33 frac-
tured reservoirs. Heffer et al., 1997, Heffer, 2002. See also extensive data sets in Tingay et al. 2005.

rosettes), that the unloading caused by the two water- The new sub-30 Hz data was shown to suffer signifi-
flood effects mentioned above, are in many cases cantly less attenuation due to scattering, and revealed
stimulating shear-failure (of ‘intact’ rock), or shear deep reflections not previously seen in conventional
displacement, and therefore further dilation, of existing surveys. One of the problems identified in numerous
joint or fracture sets, or faults. previous attempts had been the scattering and attenua-
The direction of H max shows a frequent tendency to tion from the rough basalt surface, and absorption due
have bisected the geologic features that are the basis for to faults and joints, plus interference from inter-bedded
the rosettes shown in Figure 14.33. The mechanism units of e.g. claystone and siltstone.
may (often?) be one of conjugate shear, in which case Strong sea-surface reflections had also been a problem.
showing strong parallels to the findings of Townend Basalts of this region are very heterogeneous at scale
and Zoback, 2000 concerning the frequency of water lengths of tens of metres, approximately an order of
conducting features being under shear stress. These magnitude less than the seismic wavelength when
topics are addressed in detail in Chapter 16. of 10 Hz. So use of the long wave-lengths could avoid
the problems of thin inter-layering and lateral inhomo-
geneity.
14.12 Low frequencies for The authors generated synthetic seismograms, using
sub-basalt imaging an available sonic well log in basalt, and assumed a sin-
gle deep reflector. They used the following model:
Many prospective ocean margins are covered by large
areas of basalts. These tend to be extremely heteroge- 1. 0–400 m Vp  1500 m/s
neous, and scatter and attenuate the seismic energy 2. 400–1000 m Vp  2500 m/s
of conventional seismic reflection surveys, making it 3. 1000–1800 m Vp  3500–5500 m/s (strong oscil-
difficult to obtain seismic images from deeper reflect- lation in basalt flows, greater extremes)
ors. Ziolkowski et al., 2003, argued that since high fre- 4. 1800–5000 m Vp  3000 m/s (assumed sediment-
quencies were scattered more than low frequencies, it ary basin)
would be logical to emphasise low frequencies, by using 5. 5000–6000 m Vp  4000 m/s
much larger air guns, towing the source and receivers
at greater depths than usual, i.e. 15–20 m. This was The modelled basalt was 800 m thick. Their subse-
done in 2001, over an area of the NE Atlantic margin quent field survey, using large air guns towed at 15 m
that holds promise of very large accumulations of hydro- depth, were able to image deep reflectors in a much
carbons in the Mesozoic and Palaeozoic sediments, clearer manner than achieved in a conventional survey.
which are covered by higher velocity Cenozoic flood We may conclude this section with an exotic use of
basalts. reflection seismic imaging, described by Dypvik et al.,
404 Rock quality, seismic velocity, attenuation and anisotropy

1996 and Tsikalas et al., 1998, namely the imaging, in the


Barents Sea, of the 40 km diameter Mjølner impact struc-
ture, originating some 145 million years ago, which
involved an 850–1400 km3 disturbed volume of rock,
possibly caused by ‘only’ a 0.9 to 3 km diameter meteor. A
transient crater of 16 km diameter and 4.5 km depth, and
ejecta of some 175 km3, were interpreted from the seismic
images. Large amplitude tsunami waves, and energy
release of some 1020 or 1021 Joules, equivalent to an 8.3
magnitude earthquake were estimated by the authors.
Figure 14.34 Fracture-type schematic for the Clare field reservoir,
given by Smith and McGarrity, 2001.
14.13 Recent reservoir anisotropy
investigations involving
P-waves and attenuation

Mapping the azimuthal velocity anisotropy of P-waves,


to reduce risk of drilling low-productivity wells in rela-
tively unfractured parts of the UK offshore Claire Field,
west of the Shetland Islands, was described by Smith
and McGarrity, 2001. Their company, BP, had per-
formed multi-azimuth walk-away (‘float-away’) VSP in
an offshore well to attempt a calibration with oriented
core data and with FMI logs. Three 4 km walk-away lines
were performed, using a downhole triaxial accelerometer
sensor array, both in the overburden (1486–1535 m)
and in the fractured reservoir (1850–1906 m). They
described the obvious importance of ‘open conductive
fractures’ with producing-well intersection of these fea-
tures as the goal.
The authors cited azimuthal variation in the shear
modulus of the fractured rocks as the reason for the
P-wave velocity anisotropy. ‘P-waves travelling obliquely
to the fractures can be influenced by the shear proper-
ties of the rock’. This is perhaps an alternative way of Figure 14.35 Polar plots showing orientation from core, FMI frac-
acknowledging the theoretical contribution of the shear ture orientation, and maximum stress orientation
stiffness (and normal stiffness) of the fractures to the (90° to calliper-log elongation?), compared with
anisotropy, which also causes shear wave splitting, as we travel time inversion (Vp anisotropy) and polariza-
shall see in Chapter 15. tion inversion, using postulated velocity models to fit
the direction of wave front motion at VSP downhole
The authors also mentioned the more conventional
receivers. Smith and McGarrity, 2001
dominance of one fracture set orientation, with an
orthogonal subset, and mentioned variation of fracture
density in this unequal two-set system, as a reason for (almost) H-parallel ‘open’ fractures as the clear source of
variations in the degree of seismic anisotropy. We will preferred orientation for both the travel time and the
investigate fracture density in detail in the next chapter, polarization inversion, as shown in Figure 14.35. The
mostly in connection with the interpretation of shear- magnitude of the (P-wave) velocity anisotropy, typically
wave anisotropy. 6%, actually ranged from 0 to 15%, and the authors
A schematic classification of fracture types in the Claire were able to demonstrate reasonable correlation of high-
reservoir, shown in Figure 14.34, suggesting considerable flow wells with higher (P-wave) velocity anisotropy and
complexity, reduced nevertheless to the predominance of low-flow wells with lower anisotropy.
P-waves for characterising fractured reservoirs 405

The very rough correlation between velocity anisotropy


and well flow rate, are shown in Figure 14.36b. It was
therefore suggested that conventional P-wave seismic
acquired by towed arrays in marine environments may
help characterize fractured reservoirs, thereby maybe
avoiding the need for expensive multi-component shear-
wave acquisition.
Following on from the above azimuthal anisotropy of
P-waves at the Claire Field, as discussed by Smith and
McGarrity, 2001, a later study by Maultzsch et al., 2005,
with BGS colleagues from Edinburgh, addressed the
question of attenuation anisotropy at this field. They also
used the multi-azimuth walk-away VSP, and demon-
strated that the fractured, oil-saturated reservoir also
showed a consistent azimuthal variation in attenuation.
These authors first demonstrated the 1/Q seis related
(a) modelling abilities of the Chapman dynamic poroelastic
matrix-and-aligned-fractures model, which is reviewed
in Chapter 15. A simulation of P-wave attenuation in the
case of a modelled rock mass with vertical fracturing is
shown in the polar diagram reproduced in Figure 14.37,
showing in b), a maximum attenuation 1/Q  0.15 for
the vertical direction of fractures.
First they measured a zero-offset Q seis, using data
from all source offsets and azimuths. They obtained
Q seis  18 for the reservoir which was fractured (again
a very plausible match to deformation modulus in
GPa), and Qseis  35 to 40 for the overburden, which
they assumed was relatively unfractured. Secondly they
(b)
analysed attenuation as a function of azimuth for each
Figure 14.36 a) Density of shading represents degree of P-wave offset, both in the reservoir and in the overburden, util-
velocity anisotropy at reservoir (unit V) level. b) Well ising two tool settings. In the reservoir, minimum
flow rates versus estimated P-wave velocity anisotropy. attenuation lay between N70°E and N100°E. But at
Smith and McGarrity, 2001. the shallower receiver setting in the overburden the

(a) (b)

Figure 14.37 Two examples of attenuation modelling with a set of vertical aligned fractures, using Chapman’s dynamic poroelastic matrix-
and-fracture-set model. See Chapter 15 for a description of this model. (Reproduced by kind permission, Maultzsch pers.
comm. 2005). (see Plate 5).
406 Rock quality, seismic velocity, attenuation and anisotropy

Figure 14.39 A sheared, rough fracture analogue for ‘open’ frac-


tures, that shows the dilated parts (white) as having a
different average orientation compared to the black,
contacting (partly crushed, load-bearing) parts of the
asperities. Barton, 1973.
Figure 14.38 a) Fracture rose diagrams for all fractures, and b) open
fractures at the Claire Field. c) measured minimum
attenuation direction from multi-azimuth walk-away
VSP. d) major horizontal stress estimate. (Maultzsch using hydraulic fracturing, one must question whether
et al., 2005, and Smith and McGarrity, 2001). there can be ‘open’ fractures in a petroleum reservoir at
right angles to h min, unless:

minimum attenuation was scattered between wider a) the rock is unusually strong and that joints are
azimuths, as shown in Figure 14.38a. rough,
Rose diagrams of all the fractures and of the open con- b) there exists a close-to-fracturing pore pressure,
ducting fractures, from others’ logging of cores and bore- c) that shearing has occurred,
hole images are reproduced in Figure 14.38a and b. d) there is a suitable quantity of hard mineralization
Maultzsch et al., 2005, also provide, in Figure 14.38c, to ‘bridge’ and maintain an earlier porosity.
their measured azimuth of minimum attenuation, which
although close to a match with the open fractures, is Fractures can presumably be conceived as ‘open’ if
actually some 20° oblique to these. Curious to see where they have good connectivity and reasonable apertures
the major horizontal stress is oriented, we can add this as (for example 10 to 100
m conducting apertures?)
in Figure 14.38d, from Smith and McGarrity, 2001, just with respect to the gradients that are operating.
reviewed. It is close to parallel to the open fractures, yet If limited shearing and therefore dilation has occurred
the minimum attenuation was 20° different. in the past, and if there are limited influences from a sec-
A possible explanation is that the ‘open fractures’ are ond set of fractures, then Figure 14.39 can be a possible
open due to limited (but sufficient) shear dilation explanation for the apparent 20° rotation in the above
episodes in the past, when H max was perhaps not par- case. Note the possible influence of a rotation of the
allel to these joints. From a rock mechanics point of fluid-bearing parts of the fractures (black), in relation to
view, and armed with several experiences of measuring the contacting parts taking the load. This model will be
a considerable h min magnitude at reservoir depths developed further in the last two chapters of this book.
15 Shear wave splitting in
fractured reservoirs and
resulting from earthquakes

In this chapter the effect of structural anisotropy on rock joints will be referred to in several contexts, where
shear wave splitting and polarization phenomena will be deemed appropriate. There will also be a strong focus
treated in some detail, due to its extreme importance on the possible links between the joint or fracture shear
in helping to characterize jointed or fractured reser- and normal compliance used by geophysicists, and the
voirs, and due to the improved insight it is giving into macro-deformation, and inverted stiffnesses used for
earthquake phenomena. The structural anisotropy may many years in rock mechanics models of jointed media.
be stress aligned, and there are then logical ties to the The need for compatible measures of volume-defined frac-
principal permeability or drainage directions. Dominant ture densities and in situ values of compliance (as opposed
jointing and natural fractures are of increasing interest to those obtained from hand-sized joint samples or
to petroleum companies, both for production and for aid- roughened plates of Lucite) is necessary for further
ing stimulation, where matrix permeability is low but development.
hydrocarbon storage high. This vitally important struc-
tural feature is notoriously poorly sampled by vertical
core and well-bore scanning, since itself often sub-vertical 15.1 Introduction
or vertical. The ‘miracle’ of shear-wave splitting (with
assistance from azimuthal AVO P-wave surveys) has In view of the very widely accepted knowledge that ver-
provided the means of detecting the presence of these tical and sub-vertical jointing is extremely common in
compliant, fluid-bearing fractures. Contrary to the clas- most rock masses (due to such diverse effects as cooling,
sic wisdom of porous media fluid substitution theory, bed-flexure and tectonic influences of horizontal stress),
in situ fractures also seem capable of signalling to the it is unfortunate to say the least, that vertical boreholes
shear-waves, whether they contain gas or brine, through are usually the first, and seemingly also the second choice,
subtle velocity reduction of the slow S-wave, due to the for sampling and gaining access to the sub-surface. As
fluid-compressibility-altered fracture compliance mag- pointed out by numerous authors, and also quantified
nitudes. Geophysicists utilise an unfortunately ambigu- by R. Terzaghi, 1965, the sampling bias caused by the
ous way of describing fracture density: as number per unit mismatch of borehole diameter and horizontal spacing
volume times radius cubed. This ambiguity, meaning of vertical structure, and the vertical borehole itself, is
that millions of microcracks or a hand-full of fractures extreme.
can give the same magnitude, nevertheless seems to have Our vertical boreholes provide such a poor sample of
a remarkable proportionality to shear-wave anisotropy, the jointed sub-surface, that P-wave azimuthal anisotropy
but clearly this can be altered by changed compliances. on the one hand, and shear-wave splitting caused by the
The need to define a specific volume of fractures for presence of vertical or aligned structure on the other
reservoir understanding is urgent. Consequently, recent hand, are truly god-given means for rectifying our poor
numerical dual-porosity poro-elastic modelling devel- sampling strategies. If the ‘economy’ of a vertical well as
opments have become increasingly important for explor- opposed to a steeply inclined one, and the subsequent
ing the frequency-dependent velocity and attenuation ‘cost’ of an extensive seismic survey were combined, there
resulting from the various potential scales of anisotropy. would perhaps be more reason for rapidly deviating our
Case records both from seismology and petroleum engin- boreholes at least 10° or 15°, in order to sample the
eering will be used to show recent trends in analysis increasingly understood relevance of vertical and sub-
of seismic survey data. Relevant rock mechanics experi- vertical structure on hydrocarbon production. As empha-
ence with coupled stress-deformation-flow testing of sised by Laubach, 2003, a central challenge of sub-surface
408 Rock quality, seismic velocity, attenuation and anisotropy

fracture characterization is obtaining data on essential Shear waves were once considered just noise and had
fracture attributes where direct observation is (remark- to be filtered out. Now, with proper multicomponent
ably) unlikely. recording, S-waves can deliver important information
Recently, five boreholes inclined just 30° from the hori- both concerning rock and fracture properties, and con-
zontal, with a combined length of 1.4 km were used to cerning fluid type in the case of shear-wave splitting.
investigate the steeply folded geology and structure along
the route of a rail tunnel in Norway. The value and rele-
vance of the core logging and permeability measurements 15.2 Shear wave splitting and
was thereby increased 10- or 100-fold in relation to the its many implications
equivalent length of vertical holes. Of course this is not
a recipe for deep oil-field wells, but a fairly early devia- A landmark paper on the relatively new technique of
tion of 10° or more would greatly improve understand- shear wave splitting, summarising the first ten years
ing of both the overburden jointing and the reservoir of developments, was given by Crampin and Lovell,
jointing. The potential anisotropy of the overburden 1991. Following theoretical developments of Keith and
cannot be ignored in seismic inversion. Crampin, 1977, and the suggestions of Crampin, 1978,
As Barkved et al., 2004 pointed out in their review of shear wave splitting was positively identified above small
multicomponent data, countless reservoirs have been earthquakes by Crampin et al., 1980, and many times
discovered, characterized, and monitored by P-waves. subsequently. Shear wave splitting had already been
However, P-waves cannot solve every seismic imaging noted in a number of sedimentary hydrocarbon basins,
or reservoir description problem. With the addition of with Lynn and Thomsen, 1986, and Willis et al., 1986,
S-waves, (usually in the form of converted PS-waves, see reporting at the 56th SEG (Society of Exploration
Chapter 14), oil and gas companies have found an enor- Geologists) meeting in Houston.
mous quantity of new reserves that could not have been A simple schematic of the shear wave splitting prin-
found with P-waves alone. The new reserves have been ciples, given by Crampin and Lovell, 1991, is reproduced
more effectively exploited by better identification of in Figure 15.1a. When a steeply inclined shear wave
fracturing, and therefore better placement and devia- meets sub-vertical, aligned discontinuities, the shear wave
tion of the wells. Not infrequently, shear-wave technol- splits into two components, which have different arrival
ogy provides information where shallow gas has obscured times and different, usually orthogonal polarization. Both
P-wave imaging over central parts of a field. A particu- these aspects are shown in Figure 15.1.
larly good example is the Ekofisk reservoir. A more complete version of this simple diagram, from
Shear waves bring additional knowledge to a seismic Barkved et al., 2004, is also reproduced in Figure 15.1.
study due to the different rock mass properties that are This shows, depending upon ones preference for micro-
sampled. The traditional view is that shear-wave velocity cracks or joints, the simplest source of polarized shear
remains unchanged whether a formation contains gas, oil waves: a set of vertical joints or fractures. These, like
or water. However, as we shall see, because of the effects microcracks, cause the transmitted shear-wave to split
of fluid compressibility on the normal stiffness of frac- into a fast and slow component, registered as time delay,
tures (or its inverse: compliance), the shear waves passing due to the attenuating effect of fracture compliance on
through a fractured or jointed medium, will give the abil- the S-wave component that has particle motion perpen-
ity to distinguish between oil (low compressibility) and dicular to the fracture strike. The difference in travel-
gas. (Van der Kolk et al., 2001). time between the fast and slow waves (termed qS1 and
There is possibly a small point of controversy here, qS2 elsewhere in this book), is strongly related to fracture
concerning whether the fractures and shear-waves can density, and also to fracture compliance, since both have
both be vertical, when theoretically only shear compli- an attenuating effect.
ance is sensed. It is uncertain if shear compliance will be An often referred aspect of shear wave splitting and
sufficiently affected by the fluid compressibility. A more polarization is that the faster of the two polarized com-
certain effect of fluid compressibility – i.e. the ability to ponents is parallel or sub-parallel to the direction of
distinguish between oil and gas is when normal com- maximum horizontal stress and/or to the preferentially-
pliance is involved in the case of sub-vertical fractures oriented fluid-filled microcracks, cracks or sets of sub-
(or sub-vertical shear-waves) giving a finite incidence vertical joints or fractures. The shear-wave velocity
angle (Sayers, 2002b). anisotropy is often in the range 0.5 to 5%, sometimes
Shear wave splitting in fractured reservoirs and resulting from earthquakes 409

much more. The distribution of stress-aligned cracks or


inclusions is referred to by many authors as ‘extensive-
dilatancy anisotropy’ (or EDA), following Crampin’s
publications.
Fracture (or joint) characterization using seismic
methods is considerably more complicated than for the
isotropic case, due to the existence of the three distinct
body waves, which propagate with different velocities
and polarizations. Following general convention, these
waves are referred to as the qP (quasi-compressional),
qS1 and qS2 (fast and slow quasi-shear) modes, or sim-
ply S1 and S2.
At the time of the Crampin and Lovell, 1991 review of
(a) the first ten years of recordings (which accelerated rapidly
with the development of digital, three-component geo-
phones in the late 1980s) it was still not known exactly
what caused the shear-wave splitting, because aligned
pore space, layering and consistently aligned micro-
cracks and joint (or fracture) sets could each cause the
phenomenon.
Shear-wave splitting had been identified by this time,
in a wide range of rock types. It was considered remark-
able that, with all the different scales and characters of
aligned fluid-filled cracks, inclusions or fractures in sedi-
mentary, metamorphic and igneous rocks (with their
wide differences in strength and modulus), the differ-
ential shear-wave anisotropy varied only within narrow
limits (0.5 to 5%). Since that time, with increasing appli-
cation at well-fractured reservoirs, this range has been
greatly exceeded, as we shall see later.
The magnitude of shear-wave anisotropy appeared
already to correlate well, with the amount of hydrocarbon
production where there was larger-scale jointing or frac-
turing. There was also evidence, e.g. from Lewis, 1989,
(b)
that the delay between the split shear waves appeared to
Figure 15.1 a) Simple schematics of shear-wave splitting principles, decrease with increasing depth, yet an accumulative delay
as a direct or indirect result of the principal stress with increasing depth had been expected.
and the associated stress-aligned, fluid-filled cracks. Crampin and Lovell identified several difficulties with
Crampin and Lovell, 1991. An updated version of this shear wave splitting interpretation in the case of earth-
classic diagram, from the BGS Anisotropy Project in quakes. Because of the relative steepness needed for the
Edinburgh, shows larger ‘cracks’, signifying not just incident wave to make an acute angle to (typical) sub-
aligned microcracks, but perhaps an aligned joint set, vertical structure, there was a need for the recording site
also assumed to be roughly parallel to maximum stress.
to be within a so-called ‘shear wave window’. The epi-
b) A more comprehensive diagram of the principles of
central distance from the recording sites needed there-
shear-wave splitting, from Barkved et al., 2004. The
fast qS1 particle motion is polarized in the average
fore to be considerably less than the focal depth of the
direction of fracture strike, while the slow qS2 particle earthquake.
motion is polarized perpendicular to the average frac- Another limitation with interpretation of shear wave
ture strike. Note the ‘formation fast axis’ substitution splitting from earthquakes was the combination of the
for major horizontal stress. They may of course be large scale and the generally complex geology and tec-
synonymous. tonic structure between the source and the recordings.
410 Rock quality, seismic velocity, attenuation and anisotropy

The shear waves would likely pass through a range of rock as we shall see in Chapter 16, there are other joint or
types with different ages. Velocity, fracturing style and fracture alignments (caused by shearing), that are of
individual fracture-set properties would therefore likely major importance too.
vary, and each split shear wave could therefore split again, Temporal changes of shear wave splitting were referred
giving multiple splitting, with the influence of the joint to already in Crampin and Lovell’s ‘10 year review’ paper.
structures near the recording site as one of the prominent These had occurred before and after earthquakes, and
results. Multiple splitting obviously makes time delay after hydraulic pumping in granite. Changes in the aper-
estimation more difficult. ture of the fluid-filled, aligned features (cracks, joints or
Crampin and Lovell emphasised that the fluid filled fractures) would presumably have changed the aspect
features (cracks, inclusions, fractures) being the most com- ratio of these features, and changes in Kn/Ks ratios would
pliant (i.e. with least stiffness, as referred in rock mechan- be a macro-deformation rock mechanics consequence
ics), would potentially be most sensitive to pore pressure of such changes.
changes or deformation, and thereby modify the way that In the case of earthquake source investigations, it is
the shear waves pass through the rock mass. unfortunate that shear-wave polarizations tend to be
determined by the anisotropic joint or fracture struc-
ture close to the recorder, which of geometric necessity
is likely to be much nearer the surface than the source.
15.2.1 Some sources of shear-
However, in the petroleum industry with its more limited
wave splitting
depths, down-hole recording as in VSP, can focus on
the structural domain of interest. For this reason, many of
Crampin and Lovell, 1991, listed five possible scales
the important developments up to the time of Crampin
of azimuthal anisotropy that could cause shear wave
and Lovell’s 1991 ‘10-year’ review, had been made with
splitting:
VSP configurations.
Two further important practical details concerning the
1) aligned crystals potential uses of shear-wave splitting technology were
2) direct stress induced anisotropy emphasised by Crampin and Lovell, 1991. Firstly that
3) lithological anisotropy (e.g. aligned grains) the shear-wave-train likely contained many times the
4) structural anisotropy (fine layering) information carried by the P-wave-train, and that this
5) stress-aligned crack-induced anisotropy was in the wave forms themselves rather than just in
arrival times. Secondly that the multiplicity of source-
Subsequently, much evidence for larger-scale fracture- to-geophone ray paths required to analyse P-wave arrival
or joint-set alignment effects on shear wave splitting have times, was not needed to obtain information from shear
been obtained, which is a very important ‘addition’ for waves.
shear wave splitting, in view of the importance of these The problem of relative scale was also emphasised by
structural alignments for hydrocarbon production from Crampin and Lovell, 1991. The likely dimensions of the
fractured (naturally jointed) reservoirs. The larger scale different phenomena (from extension-dilation cracks
features dominate drainage potential from the matrix to perhaps of microns to fractions of millimetre size, to frac-
the joints, and thence to the wells, while the smaller scale tured reservoir features of many metres size) were each
EDA(extensive dilatancy anisotropy), popular with much smaller than the wave-lengths of most shear waves
Crampin in many subsequent papers, would logically (i.e. tens of metres in reflection experiments to several
dominate drainage from the pores to the microcracks. kilometres in teleseismic shear waves). It was emphasised
However, Crampin and Lovell emphasised that it was that multi-offset, multi-azimuth, three-component VSP
the stress-aligned vertical fluid-filled cracks that were was likely to be the best way to attempt to analyse the
most likely to cause more uniform splitting. That the geometrical aspects. Some of these investigations will be
two shear wave components will be oriented (polarized) reviewed later.
parallel (or sub-parallel), and perpendicular (or sub- In Chapter 10 concerning the phenomenon of atten-
perpendicular) to the present most permeable fracture uation, the inverse trends of velocity and attenuation
set directions, that may themselves be parallel (or sub- magnitudes were seen. High velocities only occur where
parallel) to the present major horizontal stress is the logical attenuation is low or Qseis is high. Links to rock mass
extension from crack-scale to joint or fracture scale. But qualities Q and Q c through deformation moduli were
Shear wave splitting in fractured reservoirs and resulting from earthquakes 411

also cited, i.e. high velocities only occur where rock qual- Let us examine some orders of magnitudes to get a ‘feel’
ities Q and Q c are high. Q seis was likened, numerically, for this actually rather ambiguous geophysics parameter:
to the rock engineering static deformation modulus in
the first kilometre (when expressed in GPa), and to the 1) Microcracks: Assume 106 microcracks of radius 100
Q c value itself at extreme depth, since both seem to microns in a 10  10  10 cm rock sample. This
depend on the degree of jointing, clay-content, and rock gives e  106  (104)3/(101)3  0.001. If the
hardness, despite the micro-displacements involved in ‘microcracks’ were of 1 mm size, the crack density
seismic loading. would be 1.0. The ‘crustal range’ is contained within
Crampin and Lovell made the interesting observa- these ‘extremes’. (The microcracks would need to be
tion that the reciprocal relationship between velocity 330 to 465
m to lie in the mid-range of e  0.01
and attenuation (e.g. Crampin, 1981) was one of the to 0.05) with the above frequency.
reasons why the leading split shear wave was such a sta- 2) Fractures (or joints): Assume 10 joints of 1 m radius
ble phenomenon; because it was travelling in the fast in a 10 m3 block of rock in situ. We have: e  10 
direction and was less attenuated than the slower split 13/10  1.0. (This represents a well-connected
shear wave. The oriented rock mass quality Q o described rock mass if for instance there were on average 3.3
by Barton, 2002, also has its maximum value in this joints in each principal direction of the 2.15 m ‘cube’.
same direction, i.e. sub-parallel to dominant structure, If there was only one aligned set, the spacing would
due mainly to higher RQD in this direction. (The use be close to 20 cm, if two conjugate sets, the spacing
of oriented RQD, termed RQDo, gives Q the directional of each would approximate 40 cm. This fracture
dependency Q o, together with the changed Jr/Ja ratio density is significantly less than much of the jointed
representing the frictional strength of the joints across chalk at Ekofisk, as we shall see later. If there was
the sampling direction. See Appendix A for description only one such joint per 1000 m3 (10 m on a side),
of these terms). the ‘crack density would be the same as our more
Crampin and Lovell, 1991, concluded their ‘10-year extreme microcrack example with e  0.001. (We
review’ by claiming that ‘progress in understanding shear- would need 10 to 50 joints of 1 m radius per 10 m
wave propagation is the most fundamental advance in cube to lie in the mid-range of e  0.01 to 0.05).
seismology for some decades’. They also speculated on 3) Fault swarm: Assume 10 medium-sized faults of
the future uses of shear-wave splitting, including moni- 464.5 m radius in 1 km3. In this case e  10 
toring of hydrocarbon production and monitoring the 464.53/109  1.0. If we veer to the other extreme,
stability of major civil engineering works in rock. The we would require only one minor fault of 100 m
former is now much used. radius in 1 km3 to give e  0.001. (We would need
10 to 50 minor 100 m faults per 1 km3 to lie in the
mid-range of e  0.01 to 0.05)
15.3 Crack density and EDA
To a non-geophysicist, it is difficult at first viewing,
The geophysicist’s crack density is defined as: to see why so much seismic interpretation is related to
this parameter when it is so ambiguous. To check again:
Na 3 (15.1) 1) microcracks. e  107  (104)3/(101)3  0.01 (ten
e
V million @ 100 ␮m/10 cm cube)
2) fracture e  10  13/103  0.01 (ten @ 1 m/10 m
where a is the crack radius, cubed due to the argument cube)
that this relates to the energy of elastic deformation asso- 3) minor fault e  10  1003/10003  0.01 (ten @
ciated with the crack. (O’Connell and Budansky, 1974). 100 m/1 km3)
N is the number of such cracks in volume V. According
to the authors Leary et al., 1990, crack density is often These three very unequal scenarios with their equal
in the range of e  0.01 to 0.05 in widely different geo- ‘crack density’ would inevitably have totally different
logical and tectonic regions. This opinion seems to stem mechanical and fluid-conducting properties. Yet sur-
from the articles of Crampin, several of which will be prisingly, they are supposed to generate equal shear wave
reviewed here. anisotropy, as we shall see.
412 Rock quality, seismic velocity, attenuation and anisotropy

The Leary et al., 1990 review of the 1988 Chapman focussed on this scale of EDA for many years, while
conference (and 3rd International Workshop on Seismic others have now demonstrated dispersive effects caused
Anisotropy) papers on ‘Seismic Anisotropy of the Earth’s by fractures of entirely different dimensions, using dou-
Crust’, gave a useful broad-brush reflection of some of the ble-porosity or triple-porosity poroelastic models, with
earlier opinions on fracture and crack induced anisotropy pore-microcrack-fracture-fluid interaction.
and its measurement, prior to the great advances in data The ‘problem’ is that the delayed, approximately
acquisition and computer processing, that occurred orthogonal shear-wave arrivals are also consistent with
mostly during the 1990s, that has given ‘anisotropy’ a cen- a feasible density of aligned vertical microcracks. How-
tral role in earthquake interpretation and fractured reser- ever, the relative stiffness of microcracks, having much
voir exploration and subsequent production monitoring. higher aspect ratios than inter-locked fractures or joints,
Aligned fracturing may be detected and monitored means that they cannot respond in the same way as frac-
over a huge range of length scales, using polarized shear tures, to a given change in fluid pressure – at least
waves. Dimensions may range from crustal dimensions according to classic geophysics teaching. Furthermore
of 10–100 km, through 1–1000 m reservoir scale frac- there is the crack density definition, which is a source of
tures and faulting, to millimetre and micron-sized micro- confusion (for non-geophysicists).
cracks. The potential of fluid-filled microcracks to react
to crustal stress and strain led Crampin et al., 1984 to
propose the ‘extensive-dilatancy-anisotropy’ (EDA) con- 15.3.1 A discussion of ‘criticality’
cept. The EDA hypothesis is that crustal fluids prop due to microcracks
open a population of high-compliance voids or inclu-
sions that are nevertheless capable of remaining open According to Crampin, 1993b, several oil companies
against the least principal stress. The implication of 3D were already reporting shear wave splitting in almost all
principal stress anisotropy at depth is that EDA cracks their three-component reflection surveys in sedimentary
will tend to be aligned in a vertical plane, striking par- basins. Furthermore, perhaps contrary to Crampin’s
allel to the major horizontal stress. With this configura- expectations, the splitting, as he observed, was assumed
tion, Crampin 1978 had reasoned that a microcracked to be due to large fractures within fractured reservoirs.
but otherwise isotropic crust was transversely isotropic, Shear wave splitting was also visible in reflections from
with a horizontal symmetry axis. layers above the reservoirs, apparently suggesting to
Shear waves polarized parallel to the microcracks, Crampin that exclusive dependence of splitting on large
and also travelling parallel to them, hardly sense the fractures seemed unlikely.
presence of the cracks, and travel at almost the wave In fact as we will see later in this chapter, shear wave
speed of the unfractured matrix. However, the shear splitting is also seen to follow the saucer-shaped subsid-
waves travelling in this same parallel-to-structure direc- ence bowl far above a compacting North Sea jointed
tion, with polarizations normal to the microcracks, chalk reservoir, with correlation to the exact location of
sense the reduced shear stiffness caused by the cracks increased sea depths. Joint-stretch in the overburden
and are slowed, in a similar manner to shear waves that seems likely to be the cause, but if unconsolidated sedi-
might be travelling along the (perpendicular) axis of sym- ments were the actual source of the splitting/polariza-
metry. As a result of the differential wave speeds, the tion, microcracks or even macro-cracks in the sediments
shear waves travelling parallel to the aligned microfrac- would need to be invoked to explain the polarization
tures, with their two different polarizations, separate in match to the subsidence bowl.
time in proportion to the length of travel path and the It is not quite clear why, but Crampin, 1993a was of
density of the crack population. The majority of polar- the opinion that fluid-filled microcracks were the most
ized shear wave observations show the fast wave polar- compliant elements of the rock mass. If this opinion
ized parallel or sub-parallel to the accepted local or was because of the assumption that larger scale (and
regional maximum stress field. lower aspect-ratio) fractures and joints would be closed
EDA can in principle refer to a wide range of crack at depth, then indeed the usually less compliant microc-
sizes, including fluid-conducting fractures. However, racks could remain as perhaps the most compliant ele-
it seems that because many of the observed seismic ment of the rock mass. But here we run into difficulties
properties can be simulated by propagation through concerning the interpretation of 4D repeated surveys
distributions of microcracks, Crampin appears to have over producing reservoirs.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 413

Rock physics provides evidence of effective stress sen- pavements, where the pavements are characterized by mas-
sitivity of e.g. S-waves, P- and S-wave attenuation, and sive jointing and the rock cannot be considered as intact.’
of their anisotropy and dispersive nature. (Chapter 13).
In a later article, Crampin, 2000 relates e  0.055 to
When changes of oil saturation in producing parts of a
‘fracture criticality’, and states that ‘almost all rocks are
reservoir are monitored, or water/oil front progression
(then) marginally close to the ‘critical percolation thresh-
is monitored (see examples in Chapter 14), it is surely
old in a stressed fluid-saturated solid, when shear strength
the fractures that are the primary targets for the diagnos-
is lost and fracturing occurs’. It has to be emphasised that
tic seismic waves. Do the pore-to-microcrack responses
these are foreign concepts to rock engineers: extensively
to production-caused pore pressure, occur fast enough to
jointed rock masses with far higher crack densities have
be ‘in-step’ with the responses in the probably more com-
perfectly adequate strength to tolerate e.g. tunnelling.
pliant fractures and joints? After all, the fractures have
Rock stress and joint roughness helps retain both
far higher permeability in general, than the matrix. As we
strength and permeability. If there are two or three joint
have seen in the earlier critique of ‘crack density’, this
sets contributing to ‘e  0.055’, there is actually little
parameter does not seem to help in answering this ques-
cause for concern about ‘loss of shear strength’, or ‘dis-
tion, because of its basic ambiguity.
persion of fluid’. Three joints of 52 cm diameter, inter-
In his paper ‘Arguments for EDA’, Crampin, 1993a
secting (or avoiding each other) in a 1 m3 block of rock
also addressed the meaning of the typical 1% to 5% dif-
also gives e  0.055. This would be a rather stable rock
ferential shear-wave anisotropy, reportedly measured in
mass compared to near-surface experiences. When blast-
a wide range of rock types, but as we shall see, there are
ing one would see most of the remaining ‘half-pipes’ of
many exceptions in fractured areas. Crampin considered
the blast holes, but with occasional ‘block-corner’ over-
that the 1% to 5% was also equivalent to the generally
break beyond the half-pipes.
limited range of effective crack densities, which he
If the same 1 m3 block was (almost) divided into eight
assumed were usually 0.01  e  0.05. Crampin noted
component smaller cubes by three typical, near-surface,
that the percentage of differential shear wave anisotropy
mutually perpendicular joints, each of 1 m diameter, the
was usually about e  100, for a Vp/Vs ratio of about
crack density would have increased to: e  3  0.53/1 
1.7 (1.732 was quoted).
0.375. We regularly construct tunnels in such rock, and
Crampin, 1993a avoided mentioning specific rock-vol-
if stress and joint roughness are adequate, and there is
ume dimensions regarding crack density (e) in the follow-
no clay, there may be no need for immediate rock sup-
ing paragraph, which will be quoted in full, as it reveals
port of the exposed perimeter of such a tunnel. That
an unexpected way of viewing rock mass stability and the
the tunnel may leak is a different problem. It does not
fluid-bearing nature of rock masses, which is difficult to
signal ‘criticality’, rather ‘normality’. In a reservoir,
match with rock mechanics and hydraulics experiences.
‘leakage’ is desirable when in the right direction.
After all, where will the fluid go, and where will the
‘A crack density of e  0.05 (shear-wave anisotropy
rock blocks go when surrounded by neighbouring fluid
about 5%) is equivalent to a crack of diameter 0.7 in each
and neighbouring rock blocks at 3 km depth? They will
unit cube. A crack density of e  0.1 has a crack diameter
continue to contribute their coupled roles in supporting
of 0.93 in each unit cube (check: e  1  0.4653/1 
the weight of the over-lying 3 km of rock, and they will
0.100), and this is clearly near the critical crack density at
continue to provide (limited) void-space for oil storage,
which an intact rock fragments, as very close cracks begin to
and allow percolation towards producing wells. There is
coalesce to form through-going fractures. Thus the upper
surely nothing critical about this, just normality.
limit of crack densities (e  0.05, with occasional excur-
sions to 0.1) is probably due to the limit of the number of
fractures for an intact rock mass to remain intact. If this 15.3.2 Temporal changes in
limit is exceeded the rock mass fragments and the pore fluid polarization in Cornwall HDR
would disperse and, once dispersed, the cracks would tend to
close and crack-healing occur which would lead to a lower An interesting phenomenon was observed by shear-wave
crack density of open cracks. The only other occasion when splitting when monitoring the Cornwall hot-dry-rock
substantially larger crack densities (up to e  0.4) have HDR project. Crampin and Booth, 1989 reportedly
been claimed for field observations are reported by Crampin found that there were 7° to 10° rotations in the polariza-
et al., (1980a) for observations on the surface of limestone tion directions as a result of deep-well injection of (cold)
414 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.2 (a) The Cornwall joint-shearing mechanism visualized by Barton, 1986 as a result of cold water injection and effective stress
reduction in a joint set not perfectly aligned parallel to the major stress, and under a differential (shear) stress. (b) An exagger-
ated diagram of (conjugate, or single-set) joint shearing, showing the possibility of opposite rotation of the rock-to-rock contacts
(R), and open (O) fluid-filled sections during shearing of the usually non-planar joints needed to give joint permeability in reser-
voirs. This shearing-mechanism could possibly be the source of polarization rotation, both here and elsewhere, where ‘deviation’
with respect to H max has been noted.

water into the jointed and (possibly) microcracked gran- mechanism. Was the dilation considered to be a normal
ite, which was also AE monitored to several kilometres expansion of the apertures, or was there sufficient,
depth, during the pumping. Batchelor and Pine, 1986. inclined, differential stress, for shearing to occur as a
They considered that this consistent change in polar- result of the pumping? The answer would have a direct
ization was a stable, reliable and significant result. effect on how the stress field was modified.
Although the title of the Crampin and Booth paper Figure 15.2a, illustrates an alternative interpretation of
suggests an ‘explanation’ that could have been consistent the Cornwall polarization rotation of 7° to 10°, referred
with the writer’s joint-shear-dilation hypothesis (Barton, to above. The jointing at depth was under some shear
1986, see below), Crampin 1993a actually suggests the stress, as it was not aligned parallel to the major horizon-
following mechanisms: ‘The results suggest that before the tal stress. With cold water injection there is both contrac-
dilation of the incipient joints by pumping, the fluid-filled tion and effective normal stress reduction, allowing for
EDA cracks are aligned parallel to the in-situ stress field. some slight shearing that is likely to be part of the source
After the joints have been dilated by pumping and the of microseismic ‘clouds’ that propagated to greater depths
stress field modified, the EDA cracks close to the joints are down the presumably dilating, slightly shearing jointing.
realigned parallel to the joints.’ (Batchelor and Pine, 1984 and Barton, 1986). The
There follows a reference to opening and closing of hypothesis that slight joint shearing can lead to polariza-
microcracks and sub-critical crack growth, that seem tion rotation is illustrated in Figure 15.2b. The details of
not to relate directly to any well explained joint-dilation this mechanism are described in Chapter 16.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 415

Crampin, 1993a emphasised many times the (report-


edly) remarkable ‘uniformity’ of the mostly 1% to 4%
(but occasionally 10%, and even 30%) shear-wave
anisotropy, reportedly from a multitude of rocks and from
a wide range of scales of investigation, i.e. near-surface:
10’s of meters, sedimentary basins: 1–2 km, mixed geol-
ogy above earthquakes: 5–15 km, and whole crust: 30 km.
Focus on the 1% to 4% range, if generally justified, is
implying limited EDA dimensions of a few microns to a
few meters.
Crack density does not need to be 0.015–0.045 to
‘explain’ a frequent shear-wave anisotropy of 1.5% to
4.5%. Almost ‘all rock’ is unlikely to be pervaded by
tens of millions of microcracks per hand-specimen – if
it was there would certainly be a ‘critical crust’. Rather
rock masses are often pervaded by one two or three
joint sets – and it is these nearer-the-surface, very com-
mon structural-geological features, that may be imprint-
ing their shear-wave splitting result on the recorded
waves.

15.3.3 A critique of Crampin’s


microcrack model
Figure 15.3 Schematic illustration of the Zatsepin and Crampin,
Because many of the observed seismic properties can be 1997 APE model. The actually 3D, assumed ‘hexagonal
simulated by propagation through distributions of micro- boundary cracks’ are vertical. Note the four values of
increasing maximum horizontal differential stress nor-
cracks, Crampin appears to have focussed on this scale of
malized to the critical value at which cracks first begin
EDA in the last decade. Others have now demonstrated
to close. Note the remarkable, assumed increases in
dispersive effects caused by fractures of many different normalized crack aspect ratios, with increased ‘SH’
dimensions, using double-porosity or triple-porosity stress. Aspect ratios were apparently chosen to give a
poroelastic models, with pore-microcrack-fracture-fluid porosity of 5%.
interaction (see later in this chapter).
The ‘problem’ is that the delayed, approximately
by larger cracks or fractures’ … … ‘the detailed geome-
orthogonal shear-wave arrivals are also consistent with
try of fluid-saturated microcracks in almost all rocks and
an (occasionally) feasible, high density of aligned vertical
reservoirs’ … … ‘the underlying reason for the calcula-
microcracks. However, the relative stiffness of microc-
bility (of APE) and predictability (1.5% to 4.5%) is
racks, having much higher aspect ratios than inter-locked
that the rock mass is so heavily (micro) cracked that it
fractures or joints, means that they cannot respond in the
can be considered as a critical system’.
same way as fractures, to a given change in fluid pressure,
despite Crampin’s apparent claims to the contrary.
This classic aspect ratio stiffness argument, originating 15.3.4 90°-flips in polarization
at least from Walsh, 1966 seems to be discounted or
ignored by Crampin. He appears to overlook the ambigu- Figure 15.3 illustrates the anisotropic, poro-elastic or APE
ity of crack density, and trust in the assumed attributes of model of Zatsepin and Crampin, 1997, that appears to
the physically somewhat unlikely APE model reproduced have featured more and more in the interpretations and
in Figure 15.3, when making statements such as: ‘fluid- conclusions found in Crampin’s numerous recent articles
saturated microcracks are highly compliant’… …‘large concerning analysis of shear-wave splitting in both earth-
cracks would be stiff and much less compliant’ … … quake and reservoir scenarios. The dramatic increases in
‘there is very little evidence that the splitting is caused aspect ratio, the second stiffest component of rock porosity
416 Rock quality, seismic velocity, attenuation and anisotropy

(after equant porosity, and before joint or fracture poros- The ‘90°-flips’ explanation shown in Figure 15.4 sup-
ity) is supposedly able to develop in any rock, presumably poses that the assumed high (actually extreme, and surely
because of the over-looked resistance of the intervening impossible?) pore pressure causes the faster split shear-
crystals/grains of perhaps 50 to 150 GPa Young’s modu- waves that were previously parallel to H max to do a
lus, that obviously will resist such assumed expansion. 90°-flip and become the slower wave parallel to h min.
How could such a lateral volume increase, and there- It is not explained, nor is it clear in any way, how a pore
fore stress increase, from countless trillions of expanding, pressure can build up to beyond h min without dissipa-
aligned microcracks be absorbed in a rock mass without tion due to hydraulic fracturing, nor how the pore pres-
a general reversal of the H max and h min directions? sure could thereafter even approach H max.
Surely an occasional mechanism such as this is the A possible 90°-flip would appear to be possible on
actual cause of ‘90°-flips’ in the polarization, as occa- occasion with extreme H max loading, causing the usually
sionally observed/suspected in earthquake source zones. unlikely and untenable lateral expansion of aligned
(Crampin et al., 2002, Crampin and Peacock, 2005). microcracks, as depicted in Figure 15.3. What seems to
The hypothesised extreme fluid pressures that sup- be a geotechnically very unlikely process (the EPA model)
posedly can exceed h min, and even approach H max in perhaps has merit in just such earthquake-related cases
magnitude (Figure 15.4, after Crampin et al., 2002), are of ‘axial-overload’.
surely geotechnical impossibilities? If the conceptual,
hexagonal, vertical microcracks of the APE model (Figure
15.3) were miraculously present in e.g. a soft mudstone, 15.4 Theory relating joint
then, and only then, could such an aspect-ratio expansion compliances with shear
be envisaged, with or without the help of high pore pres- wave splitting
sure. They would be likely to cause a ‘90°-flip in polariza-
tion’, due to the resistance of the surrounding material. The geophysicists’ progression from seismic propagation
in isotropic media, to anisotropic layered media to trans-
versely isotropic layered media containing one set of
vertical fractures (a joint set), later increased to two sets
of perpendicular fractures, and subsequently to non-
orthogonal vertical sets, and later still to non-vertical sets,
has resulted in a progression of theoretical papers in the
geophysics literature, containing, inevitably, an increasing
content and complexity of 6  6 compliance and stiff-
ness matrices.
Here we will summarize an intermediate stage in this
progression, from Schoenberg and Sayers, 1995, by way
of introducing, as simply as possible, some necessary
theoretical aspects. These authors addressed the problem
of how the presence of a single joint or fracture set would
affect the elastic moduli of the fractured rock, since it is
the elastic moduli and the density that determine the
behaviour of seismic waves, assuming a linear, loss-free,
elastic behaviour.
Figure 15.4 Crampin et al., 2002 explanations for ‘90°-flips’
In simplest possible terms, appropriate to the non-
in polarization, as apparently observed above small
earthquakes in Iceland. Their model supposes high
mathematical treatment in this book, the effective elastic
fluid-pressures within 1 or 2 MPa of a ‘critical stress’. compliance tensor Sijkl of a rock containing fractures, is
This figure shows modelled variations of shear-wave designed to relate the average strain ij over a representa-
anisotropy with increasing fluid pressure, for five dif- tive volume V, to the average stress components ij. One
ferent sets of principal axes of stress. They presume that therefore can write:
anisotropy becomes negative for pore-fluid pressures
close to H max. (Inset shows five assumptions for ‘sh,
sH, and sv’). ij  Sijkl ij (15.2)
Shear wave splitting in fractured reservoirs and resulting from earthquakes 417

The compliance matrix that includes the effect of a set expressed as (ZT
b)/(1  ZT
b). In other words there
of vertical fractures, has to include the excess compliances is no ZN term, which requires dipping fractures, or
caused by the additional presence of the fractures. The non-vertical wave propagation for ZN to be mobilized.
authors assumed that this could be simply expressed as (In the compliance matrix the S55 term was simply
the sum (S) of the compliance of the isotropic back- 1/
b  ZT).
ground rock (Sb) and the excess compliance matrix asso-
ciated with the fractures (Sf ). The latter is composed of
the effects of a fracture-normal compliance ZN, and of 15.4.1 An unrealistic rock simulant
a fracture-shear compliance ZT related with linear slip suggests equality between
deformation. Based on the assumptions made, the single ZN and ZT
set of vertical fractures produces an unusually simple
compliance matrix: Schoenberg and Sayers, 1995 followed their vertical frac-
ture set compliance matrix treatment, by citing the results
of shear compliance (ZT) and normal compliance (ZN )
 ZN 0 0 0 0 0  interpretation from loaded, roughened Lucite (‘Plexiglas’)
 0 0 0 0 0 0  plates used ‘to simulate a fractured medium’, reported
 
S(f)   0 
0 0 0 0 0 0 (15.3) by Hsu and Schoenberg, 1993. Schoenberg and Sayers
0 0 0 0 0 
  apparently derive their assumption that ZN  ZT for
 1 0 0 0 ZT 0  the case of dry (gas saturated) cracks from this labora-
 0 0 0 0 0 ZT 
tory work.
(It is argued in Chapter 16 that interlocked rock joints
usually have quite different magnitudes of normal and
From this matrix, Schoenberg and Sayers, 1995 shear stiffness, with the expectation that at least some
developed the 6  6 compliance matrix (S) for the com- of these differences remain when considering micro-
bined transversely isotropic fractured medium, involving displacement and the (presumed) elastic response of
the various combinations of Lamé constants ( and
) joints to dynamic wave loading. However, this is at pre-
shown in Chapters 1 and 14. The simple addition of sent a subject under debate and therefore controversial).
the three fracture compliance terms (ZN, ZT and ZT) Hsu and Schoenberg, 1993 interpreted the normal and
was made in the same (1,1 5,5 and 6,6) locations in the shear compliances from the results of static and dynamic
combined compliance matrix. They then inverted this loading tests on numerous (60 or 200), roughened Lucite
to form the elastic stiffness matrix to represent the com- (‘Plexiglas’) plates, loaded normally. Their extremely close
bined isotropic back-ground and vertically fractured spacing (0.7 mm) and their continuity means that frac-
medium. They noted that this stiffness matrix was ture densities were actually in the ‘impossible’ range of
of exactly the same form as that developed previously by 7.5 to 25, up to three orders of magnitude too high to
Crampin, 1984 for the case of an isotropic medium represent fractured (or microcracked) rock.
containing a vertical array of ‘penny-shaped’ fractures, Table 15.1 indicates that these artificial-material tests
although Crampin also had second-order terms for frac- were taken to quite high normal stress levels in relation
ture density. to the assumed strength of Lucite. One can imagine
Of importance to the shear-wave splitting context of this that under the highest loads, something approaching a
chapter, are the stiffness matrix terms relating to the
fast shear wave propagating parallel to the fractures, Table 15.1 Shear and normal compliance differences obtained from
given as the C44 term, and the slow shear wave propa- normally-loaded Lucite plates with rough surfaces.
gating perpendicular to the fractures, given by the C55 Schoenberg and Sayers, 1995. Derived from Hsu and
term. For the case of the vertically propagating waves Schoenberg, 1993.
through the vertical fractures, there is no fracture com- Stress (MPa) (ZT  ZN)/(ZT  ZN)
pliance term in C44, only the Lamé constant
or
b for
the background rock. (This was inverted in the case of 6 0.1736
S44 in the compliance matrix, as 1/
or 1/
b). 12 0.1177
18 0.0332
For the case of the slow shear wave, the C55 term
24 () 0.0035
was given as
b(1-T), with T (limits: 0  T  1)
418 Rock quality, seismic velocity, attenuation and anisotropy

‘welded’ state could develop for the contacting micro-


asperities, since the ‘roughening’ would reduce the ini-
tial contact area. If a ‘locally-welded’ state developed,
this would perhaps help to explain the similar (actually
almost identical) values of ZT and ZN that they obtained.
It is recommended that the phenomena observed in
highly stressed roughened plates of Lucite, with their
extreme fracture densities, should not be used to simulate
rock joint response to dynamic or static loading, because
the suggested ‘equality’ of ZN and ZT has permeated some
of the geophysics literature, and may be far from realistic
for all but the smallest laboratory specimens.
Schoenberg and Sayers, 1995 used a ‘difference/sum’
(ZT  ZN )/(ZT  ZN ) format to compare the magni-
tudes of ZT and ZN. They used the experimental com-
ponents c11/, c33/, c44/ and c66/ from Hsu and
Schoenberg, 1993, for calculating the results given in
Table 15.1. One may ‘solve’ for specific examples of
1/Kn and 1/Ks, thus finding the ratios of difference/sum.
Let us first assume realistic small-scale pseudo-static
values of Kn  1000 MPa/mm, and Ks  10 MPa/mm,
and argue (for the sake of preliminary assessment), that
Kn  1/ZN, and Ks  1/ZT. Thus (ZT  ZN)/(ZT 
ZN)  (0.1  0.001)/(0.1  0.001)  0.98. Clearly the
low numbers for the ratio difference/sum, given in the
table, imply much closer values of the two compliances.
Suppose ZT  0.1 mm/MPa as before, while ZN 
0.05 mm/MPa. This would suggest a difference/
sum ratio 0.05/0.15, still not close enough to the Hsu
and Schoenberg 1993 highly-stressed Lucite data.
Significantly, with the increased pressure, ZT and ZN
were approaching equality.
With assumed much closer values of ZT  0.1 mm/
MPa, and ZN  0.099 mm/MPa, one obtains a ratio of
0.001/0.199  0.0050. The ‘welded’ asperity argument
for the highest stress levels would not on this basis seem
unreasonable. If approximate equality to the inverse of
stiffness was still tested, the relevant values of Kn and Ks
would be 10.1010 MPa/mm and 10.0 MPa/mm, clearly
unrealistically close in rock mechanics terms, in view
of the entirely different displacement mechanism
involved: the former involving ‘closure’ (in a stiffening Figure 15.5 Discontinuities of different scale, between sand grains,
direction), the latter involving ‘slip’ (not necessarily stiff- in a microfractured or jointed rock mass, and between
ening, possibly softening, in the absence of roughness). clay particles in shale, as depicted by geophysicists
Viewing the obviously different mechanisms involved concerned with the inequality (or unlikely equality) of
in mobilizing the ‘spring stiffnesses’ in the normal and the shear and normal compliances. S and n, or S (r)
shear directions in the various materials depicted in and n (r), are depicting the shear and normal compli-
Figure 15.5, it is hard for an experimentalist to under- ances (of the rth discontinuity) in the three or four
stand how equality could be expected between S and n, different scales of media. (Sayers, 2002a, 2002b, Liu
or ZN and ZT, since in pseudo-static testing and its et al., 2000)
Shear wave splitting in fractured reservoirs and resulting from earthquakes 419

empirical modelling, the mechanisms involved in clos- The same will apply to the fractured or jointed
ing and shearing are entirely different, and are mod- medium of much larger scale, where horizontal stress
elled in entirely different ways as a result. Some of this anisotropy will often have caused the alignment of the
difference would seem likely to remain at the micro- natural joints or fractures. (We have termed the natural
scale of dynamic wave loading. features: rock joints in most of this book, to distinguish
The question is whether heavily loaded, roughened them from artificially induced fractures, as developed
Lucite plates can be accepted as realistic models for a set around boreholes or tunnels or mine openings, or
of stressed rock fractures. Clearly they should not be fracturing induced by MHF stimulation of tight reser-
accepted, or accepted only with great caution, in view voirs. This preferred terminology is following the rec-
of the non-brittle ‘plastic’ behaviour of Lucite. It is ommendation of ISRM, 1978 and seems logical in
therefore urged that one should re-evaluate the rele- view of the number of artificial (man-made) fractures
vance of the ‘ZT /ZN  1.0’ theory, and adopt, in the that have to be described in these shared, rock engineer-
first instance, data from tests on actual rock joints, for ing disciplines.
example those described by Pyrak-Nolte et al., 1990 At the larger scale, Sayers, 2002b, deduced that ZT 
(see later), or those derived (implicitly) from Chaudry, ZN, when analysing the results of the Chaudry, 1995
1995 and King et al., 1995 polyaxial tests. The latter polyaxial tests on 41  41  41 mm cubes of sandstone,
studied principal stress-developed rock fractures in highly which had a fracture-development cycle. (See descrip-
stressed cubes of isotropic sandstone (see Chapter 13). tions of these interesting tests in Chapter 13). This
Even though both these data sets are likely of too small compliance inequality is of course consistent with the
scale, with respective dimensions of 52 mm and 41 mm, experience of Ks  Kn, concerning the pseudo-static
they are extremely likely to be closer to relevant reservoir shear and normal stiffnesses of joints and fractures, where
jointing than artificial Lucite surfaces. stiffness is the rough inverse of compliance. The mag-
nitude of Kn proves to be less than, but quite close to
1/ZN in good quality (unweathered) hard rock, while in
15.4.2 Subsequent inequality of ZN the shearing direction, Ks  1/ZT, sometimes 1/ZT
and ZT (see Chapters 13 and 16).

Later work by Sayers, 2002a, was based on elastic


anisotropy expressions involving both second-rank and 15.4.3 Off-vertical fracture dip
fourth rank tensors for describing the seismic effect of or incidence angle, and
microcrack and grain boundary contact normal and normal compliance
shear compliances in sandstones. By matching his theo-
retical predictions to polyaxial loading on blocks of Sayers, 2002a pointed out that off-axis velocity meas-
sandstones, it was found that even at pre-fracturing urements would be needed to actually quantify the mis-
microcrack scale, the shear compliance was deduced to match ZT  ZN, so he was actually not able to determine
be greater than the normal compliance, based on mis- this explicitly with the principal axes polyaxial testing
match of theory with measurement. of Chaudry, 1995 and King et al., 1995. Sayers, 2002b
We may draw some parallels between the two scales of therefore went on to address the problem of vertical
discontinuous phenomena addressed by Sayers 2002a and shear wave propagation in jointed media with off-vertical
2002b, each caused by present or past stress anisotropy. dips, using the example of two conjugate sets with
As Sayers, 2002a and other authors have pointed out, oppositely oriented dip angles, as depicted in Figure 15.6.
grain boundaries and microcracks may cause anisotropic Unlike the vertical wave propagation through the verti-
behaviour in the presence of stress anisotropy, in otherwise cal joints, with dependence of shear-wave splitting on
‘isotropic’ rock materials. The sandstone depicted only the shear compliances, as discussed earlier, the shear
in Figure 15.5a, will exhibit stress-induced P-wave wave components qS1 and qS2 depend here on both the
anisotropy and S-wave anisotropy due both to prefer- shear and normal compliances, since the incident angles
ential stress-alignment of the microcracks, and due to are no longer parallel to the joint planes.
the likely difference in the magnitudes of the normal In such cases where both shear and normal compliances
(ZN) and shear (ZT) compliances between the microc- are involved, normal compliance is reportedly reduced (i.e.
rack faces and between grain contacts. stiffened) by fluids of non-zero bulk modulus. Gas and
420 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.6 Two conjugate sets of fractures with dips of and – . Figure 15.7 A cross-discipline example of unequal conjugate joint-
This model was used by Sayers, 2002b to investigate ing used to represent a high-porosity section of the
shear wave splitting phenomena when incident shear Ekofisk reservoir, modelled by the distinct element
waves were no longer parallel to vertical fracturing. code UDEC-BB, with non-linear, effective stress and
size-dependent shear and normal stiffnesses. Barton
oil should be distinguishable by respectively greater and et al., 1986.
less shear wave anisotropy, as the stiffening effect of the
oil makes the fracture normal stiffness less contrasted to in 1985, and inspection of jointing in the Lagerdorf chalk
the back-ground medium. For dipping joints or frac- quarry in German, also supports the notion of a possible
tures, there proves to be a significant decrease in shear dominant set when conjugate jointing is present. Figure
wave anisotropy if the fluid has a higher bulk modulus, 15.7 illustrates a specific realization of jointing that was
making the normal stiffness of the fractures greater. The simulated when modelling the local-scale (1  1 m,
average of the two shear wave velocities is therefore also 2D) compaction mechanism principles, in this massive
increased. 30 km3 North Sea chalk reservoir. The behaviour of the
Bakulin et al., 2001, also pointed out that both geo- joints was based on measured values of wall strength
physical and geological data acquired over naturally (JCS) and roughness (JRC), which give pseudo-static
jointed reservoirs often revealed the presence of multi- values of both normal and shear stiffness (the latter may
ple sub-vertical fracture sets, which made the effective be significantly lower than the dynamic values). In rock
medium monoclinic. They developed a model for hand- mechanics modelling the non-linearity of these param-
ling two sets of unequal, and non-orthogonal vertical eters is also an issue, with extreme stiffening of Kn at
joint sets, which yielded the azimuths and compliances high stress, and less stiffening of Ks with stress, which in
of both sets of joints, as well as the P- and S-wave veloci- pseudo-static loading is also scale (block-size) depend-
ties of the assumed isotropic background medium. From ent (Barton and Bandis, 1982).
a second model, consisting of a single set of micro- Sayers 2002b, again claimed that the ratio of ZN/ZT
corrugated joints, they stated that monoclinic symmetry (or BN/BT in his terminology) was approximately equal
stemmed this time from coupling between the normal to 1.0 for the case of ‘gas-filled open fractures’. This was
and tangential ‘slip’. This coupling caused shear wave possibly based on Hsu and Schoenberg, 1993 results for
splitting dependence on the fluid content of the loading tests on ‘roughened plates of Lucite’, to quite
fractures. high normal stress levels in relation to the strength of the
Experience from local core logging and photography Lucite. If so, the result should be regarded with great cau-
of jointed core from the Ekofisk reservoir by the writer tion, as highly-stressed, roughened Lucite plates, with
Shear wave splitting in fractured reservoirs and resulting from earthquakes 421

three-orders of magnitude too high crack densities, surely 15.4.4 Discussion of scale effects
bear little similarity to normally-stressed, interlocked and stiffness
joints in brittle rock, as discussed earlier.
The presence of a joint-filling fluid with non-zero In Chapter 16 we will see the inequality of the pseudo-
bulk modulus in place of gas would, as suggested by static shear and normal stiffness (Ks  Kn) of virtually
Sayers, 2002b cause a lower value of ZN/ZT (or BN/BT), all joints (and filled discontinuities), and the likelihood of
but not it is suggested, just lower in relation to the changing ratios of Kn/Ks as pore pressure changes affect
above artificially high ratio of 1.0, but rather in relation the effective normal stress. These joint stiffnesses have been
to a ratio of ZN/ZT that was already significantly lower used in rock mechanics finite element modelling since
than 1.0, as seems intuitively likely due to the different the late 1960s, (Goodman and Duncan, 1968), and finite
mechanisms involved (namely ‘closure’ in the case of difference distinct element (DEM) modelling since the
ZN, and ‘slip’ in the case of ZT. (See Figure 15.5). 1970s (Cundall, 1971). The important question is
Interesting supportive field data was cited by Sayers, whether geophysicists will be able to apply the accumu-
2002b, concerning the differences in observed velocities lated knowledge from these macro-displacement,
and shear wave splitting intensity, when in the presence inversed equivalents of the geophysicists’ micro-strain
of gas as opposed to oil. (This will be reviewed later in compliances.
this chapter). We may recall that with non-vertical frac- If the stiffnesses and compliances of these two discip-
tures and with vertically propagating waves, both the lines can be related, despite the probable different orders
shear and normal compliances are assumed to be involved of magnitude of dynamic and ‘static’ deformations
in the shear wave splitting process. The field data from involved, then the more researched parameters Kn and Ks
Oman (Guest et al., 1998, van der Kolk et al., 2001) for could perhaps give shear wave splitting ‘even more infor-
a fractured (supposedly nearly vertically fractured) car- mation’ than presently assumed, since one often sees
bonate reservoir showed higher shear wave anisotropy equal normal and shear compliances assumed in papers
(due to a lower qS2) over the gas cap volume, than where describing shear wave splitting analyses, implying that
there was oil. As pointed out by van der Kolk et al., information may be limited at present, at the less access-
2001, this in situ fractured media result is in disagreement ible dynamic micro-scale of displacements.
with the prediction of Gassmann’s theory for porous Major stress-aligned, fluid-filled or sediment-filled
media, in which the shear modulus should be inde- inclusions, (and cracks, joints or fractures) are the diverse
pendent of the fluid. The shear modulus of the matrix sources of mechanical anisotropy that can be presumed
governs the fast shear wave, with no ZN or ZT influence. to be common to all rocks. They may be the guaranteed
If the in situ fractures were actually not perfectly ver- sources of pseudo-static normal to shear stiffness ratios
tical, as seems to be implied by Sayers, 2002b, then the considerably in excess of 1, most likely from about 5 to
normal compliance would also be involved in the inten- 50 as shown later in this chapter, and reinforced in
sity of the shear wave splitting. Because the normal Chapter 16. They alone actually guarantee anisotropic
compliance is reduced in the presence of a fluid with behaviour in the ‘static’ (macro-deformation or macro-
non-zero bulk modulus (i.e. oil), the ratio of ZN/ZT (or strain) regime.
BN/BT) would reduce. According to Sayers, this should The fact that the same (but inverted) units for compli-
cause a significant decrease in shear wave splitting and ance and stiffness are used by the different (‘static’ and
an increase in shear wave velocity with the increased dynamic) professions, is some guarantee that deform-
fluid bulk modulus, and therefore stiffened fractures. ation, like ‘squirt’, is taking place, however small this
Without attempting to describe complex details, it may be.
may be pointed out from recent work by Gurevich, The question that remains to be answered is whether,
2003, that the compliance matrix for a fluid-saturated, for instance, the ratio of these compliances or inverted
porous-and-fractured medium is not equivalent to the dynamic stiffnesses for application in situ, can be in any
compliance matrix of any solid medium with a single way based on dynamic tests on three joints of about
set of parallel fractures. This is due to the wave-induced 50 mm diameter, as tested in important work at Berkeley
(micro) flow of fluids between pores and fractures. (Pyrak-Nolte et al., 1990), to be reviewed next, or
Such attenuating (micro) flow is now being modelled whether there could be some hidden relation to the much
with double-porosity poro-elastic models, which are more common and easier pseudo-static tests on e.g. at
reviewed later in this chapter. least 200 samples of 100 mm length (as tested by
422 Rock quality, seismic velocity, attenuation and anisotropy

Barton and Choubey, 1977 and Bandis, 1980), or


whether in fact in situ block sizes of decimetres or meters
in size, should be the basis for in situ dynamic estima-
tion – by back-analysis.
Strictly for the case of pseudo-static loading, the
in situ block sizes were suggested by Barton and Bandis,
1982, for up-scaling (to lower in situ values), of the JRC
(roughness) and JCS (wall-strength) components of
strength and stiffness. The natural block size was sug-
gested as the limit for reducing these components, based
on pseudo-static biaxial loading tests on 400, 1000 and
4000 interlocked, fractured blocks, generated in brittle
rock-like model materials mostly consisting of dense,
weakly-cemented fine sand, as described by Barton and
Hansteen, 1979.
There seems no physical reason why a 50 mm diam-
eter core containing a joint (from a disused mine in
Sweden), should have a strong link to in situ reservoir
scale, just because this may be a convenient sampling size
also for reservoir joints. That would be fortuitous indeed.
There is clearly still less intuitive reason to use tests on
roughened plates of ‘Plexiglas’ in this extrapolation.
However, as there is so little data in this area, we must
utilise what is available, and acknowledge the following
very useful experimental contribution from Berkley in the
late eighties.

Figure 15.8 Normally loaded stress-deformation, and specific


15.5 Dynamic and static stiffness trends for two of the three jointed samples
stiffness tests on joints by tested by Pyrak-Nolte, as reported in Pyrak-Nolte
Pyrak-Nolte et al., 1990. The static and dynamic stiffnesses
for samples E 32 and E 35 are contrasted in
Table 15.2.
Several aspects of the important work of Pyrak-Nolte
et al., 1990, and of other published work with contem-
poraries at Berkeley, were reviewed in the laboratory-scale
rock physics discussions in Chapter 13. The static and when loaded to 80 MPa. The latter is remarkably stiff
dynamic normal stiffness of three natural joints, their behaviour. The contrasting behaviour possibly indicates
permeability, and the effect of normal stress on velocity, respectively ‘rough’ and more ‘planar’ joint surfaces, (or
attenuation and seismic Q were each addressed. In this differences in the degree of inter-lock), causing the
chapter on shear wave splitting, we will consider particu- widely different specific stiffness trends for these two
larly the data the authors provided on the dynamic nor- samples. (Joint roughness quantifications, and inter-
mal and shear stiffnesses. This can be inverted to compare lock descriptions seem not to have been reported).
with joint compliance data given in this chapter, and also The diameter of the jointed samples E 32 and E 35 was
in Chapter 16, where more rock mechanics pseudo- only 52 mm, which may have influenced the relatively
static data is introduced. high values of Ks(dyn), and therefore also the relatively low
The load-deformation data given by Pyrak-Nolte, values of Kn(dyn)/Ks(dyn) seen in Table 15.2. This is com-
1995 reproduced in Figure 15.8, shows the 52 mm diam- monly the case in rock mechanics pseudo-static testing of
eter jointed quartz monzonite sample E 35 (from the rock joints too: small samples have higher shear stiffness,
Stripa mine in Sweden), closing by more than 25
m, normal stiffness may be little influenced (if correctly sam-
compared to only about 5
m closure for sample E 32 pled and interlocked). Therefore the ratio of normal/shear
Shear wave splitting in fractured reservoirs and resulting from earthquakes 423

Table 15.2 Estimates of dynamic normal and dynamic shear stiffnesses derived by fitting to recorded response spectra, selected from Pyrak-
Nolte et al., 1990 data sets. Note the stiffer water-saturated behaviour. The seismic quality Qp values for these two jointed
samples are also appended. (The unit MPa/
m is suggested to aid interpretation. Note that the stiffnesses are high to very high,
with 1 MPa/
m  1000 MPa/mm).

(1012 Pa/m) or (1012 Pa/m) or


STRESS Sample E 32 MPa/
m Sample E 35 MPa/
m
n(MPa) Kn(dynamic)-dry Kn(dynamic)-saturated Kn(dynamic)-dry Kn(dynamic)-saturated

2.9 15.0 (Qp  9) 35.0 (Qp  17) 4.0 (Qp  7) 9.5 (Qp  9)
10 – 80.0 11.5 20.0
20 – (Qp  24) 100.0 (Qp  36) 20.0 (Qp  14) 25.0 (Qp  30)
70 120.0 – 32.0 59.0

STRESS Sample E 32 (1012 Pa/m) or MPa/


m Sample E 35 (1012 Pa/m) or MPa/
m
n(MPa) Ks(dynamic)-dry Ks(dynamic)-saturated Ks(dynamic)-dry Ks(dynamic)-saturated

2.9 3.5 – 1.9 –


10 9.5 – 4.8 –
20 17.0 – 6.2 –
70 55.0 – 7.4 –
See high See low
stiffness stiffness
Sample E 32 behaviour in Sample E 35 behaviour in
STRESS n(MPa) Kn(dyn)/Ks(dyn) (dry) Fig. 15.8b Kn(dyn)/Ks(dyn) (dry) Fig. 15.8b

2.9 4.3 2.1


10 – 2.4
20 – 3.2
70 2.2 4.3

Note: Qs values were generally larger than the above Qp values: see Table 10.3, Chapter 10.

Table 15.3 Kn(dyn)/Ks(dyn) ratios for sample E 30, which was of intermediate stiffness
compared to E 32 and E 35. Sample diameter was also 52 mm. Derived
from Pyrak-Nolte et al., 1990 data sets.

Normal stress (MPa) Ratio of Kn(dyn)/Ks(dyn)

1.4 2.2
2.9 1.6
6.0 1.6
10.0 1.3
20.0 2.0
33.0 2.7

stiffness tends to be low when very small sample sizes are Sayers, 1995, from dynamic tests on dry, roughened
used. There is a lot of data from near-surface rock engi- ‘Plexiglas’ plates.
neering projects where sample size is 600–1000 mm, and Dynamic stiffness ratios for a joint of intermediate
laboratory tests tend also to be in a higher 100–300 mm stiffness (E30) were also reported by Pyrak-Nolte et al.,
range compared to the high stress tests on (inevitably) 1990. This is given separately in Table 15.3.
smaller samples. Some of this data is reviewed in Chapter It may be noted from the above tables that the range
16. Note the lack of any tendency for Kn(dyn)  Ks(dyn), or of Kn(dyn)/Ks(dyn) ratios for these three small jointed sam-
its inverse ZN  ZT, as deduced by Schoenberg and ples was 1.3 to 4.3 with a mean of 2.5. This means that
424 Rock quality, seismic velocity, attenuation and anisotropy

Table 15.4 Comparison of static and dynamic normal stiffness data for the Pyrak-Nolte et al., 1987a jointed sample E 35, which was the
most deformable sample (possibly due to greater roughness). These dynamic stiffnesses were determined by curve matching.

Static stiffness Kn(static) Dynamic stiffness Kn(dyn) Ratio


Normal Stress
(MPa) (1012 Pa/m) MPa/mm (1012 Pa/m) MPa/mm Kn(dyn)/Ks(dyn)

2.9 1.0 1000 4.5 4500 4.5


10.0 2.2 2200 8.0 8000 3.6
33.0 3.3 3300 25.0 25000 7.6

the ratio of the inversed compliances ZN/ZT for these The aluminium results are therefore quite different
52 mm diameter samples ranges from 0.2 to 0.8, with a from dynamic rock joint behaviour. The actual magni-
mean of 0.4. The remaining question is what relevance tudes of dynamic normal and shear stiffness were some-
this convenient ‘core-sized’ data has to in situ reservoirs what higher than for the (obviously rougher) rock joints
in general, with a typical spread of jointed block sizes tested by Pyrak-Nolte, for instance 22.5 and
from perhaps 200 mm to 5,000 mm, and mostly sedi- 25.0 1012 Pa/m, in the case of the normal stress of
mentary rock as opposed to Stripa ‘granite’, or hard 13.2 MPa. (In alternative units these translate to 22500
quartz monzonite, which had Estatic  60 GPa, and and 25000 MPa/mm, or 22.5 and 25.0 MPa/
m, i.e.
Vp  4.9–6.1 km/s over a stress range of 84 MPa, very stiff behaviour, perhaps due to ‘asperity-weld’.
(Pyrak-Nolte et al., 1987b).
Pyrak-Nolte et al., 1987a reported an earlier set of
(dry) dynamic/static normal stiffness measurements, 15.5.1 Discussion of stiffness
performed on one of the same jointed samples (E 35). In data gaps and discipline
view of the relative scarcity of such dynamic/static data bridging needs
for rock joints in the literature, this is reproduced in
Table 15.4. In comparison with the above static and dynamics data
Units of stiffness more familiar in rock mechanics for one joint sample in igneous rock, we will see in
(MPa/mm) are also given in this table, as this unit is Chapter 16 some rock mechanics data from Bandis,
easier to visualize in practical terms (e.g. 100 MPa for 1980, with numerous static normal stiffness data for
0.1 mm closure) than that used in geophysics, where rock joints in limestone, sandstone and siltstone (and
the inverse compliance is also typically reported in ‘com- other non-sedimentary rocks). We will see Kn (static)
plex’ units of 1012 m.Pa1 (with ‘m’ sometimes omit- values varying from 250 MPa/mm at 10 MPa normal
ted altogether). It is even more ‘practical’ to consider the stress, to 31500 MPa/mm at 40 MPa normal stress.
unit MPa/
m for the first static stiffness listed in this (Priv. comm. Bandis, 2005). The lowest value was for a
table. One thereby comes much closer to the elastic, pre- rough, weathered joint (JRC  15, JCS  44 MPa) in
sumably sub-micron, dynamic wave pulse effects on limestone, the latter a smoother joint (JRC  7.6, JCS 
micro-closure, and on micro-fluid movement during 160 MPa) in almost unweathered limestone. Unstressed
squirt-flow attenuation. apertures, prior to standard stress-cycling to achieve an
In connection with the above ratios of Kn(dyn)/Ks(dyn), approximation to in situ (consolidated) conditions, were
it may be of interest to refer to the equivalent data about 0.5 and 0.2 mm respectively.
for artificial surfaces in aluminium, reported by Pyrak- Concerning shear stiffness, which is of relevance when
Nolte et al., 1992, based on interface wave experi- discussing the possible magnitude of shear compliance in
ments. An aluminium cylinder of 293 mm diameter shear wave splitting, we will see in Chapter 16 a whole
was sawn in two, planed smooth, and then sandblasted range of possible static shear stiffnesses that are seen to be
with 300
m grit. This artificial, planar-but-roughened inversely related with the sample size. The extent to
interface, which can perhaps be likened to the roughened which dynamic shear compliances are related (much more
‘Plexiglas’ surfaces studied by Hsu and Schoenberg, weakly, but perhaps directly) to sample size, seems to be
1993, showed predicted (curve-fitting-based) ratios of one of the remaining unsolved areas in this important
Kn(dyn)/Ks(dyn) ranging from 0.90 at 13.2 MPa, to 0.47 area of seismic detection of anisotropy, and the subse-
at 32.9 MPa normal stress. quent goal of interpretation of permeability.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 425

Some pieces of the ‘dynamic-permeability’ jigsaw are


‘complete’, but there are missing bridges between
dynamic and static testing, between high stress and low
stress testing, and across the related small sample to
large sample void. The ‘completed’ parts of the jigsaw
are the abilities to predict both (pseudo-static) stiff-
nesses Kn and Ks and the less tangible physical (E) and
hydraulic (e) apertures, for any size of jointed rock
blocks. This ability is based on index tests or estimates of
the rock strength (specifically the joint wall compressive
strength JCS), and the roughness (specifically the joint
roughness coefficient JRC). These parts of the ‘jigsaw’
will be illustrated in Chapter 16, in case of future use
when the three missing ‘bridges’ identified above, are Figure 15.9 Effect of joint normal stiffness on the flow rate per
robust. unit head (which is proportional to permeability.
Data for three joints in quartz monzonite, with linear
(not radial), sector-to-sector flow, crossing core of
15.5.2 Fracture stiffness and 52 mm diameter. Pyrak-Nolte et al., 1987a, Pyrak-
permeability Nolte, 1996.

Concerning one of the three ‘bridges’ discussed above, the 15.6 Normal and shear compliance
three joints in Stripa ‘granite’ (quartz monzonite) that theories for resolving fluid type
were so extensively tested and modelled by Pyrak-Nolte et
al., 1990, and Pyrak-Nolte, 1996, (see also Chapter 13 Interest in fracture-induced seismic anisotropy has rap-
review), gave interesting evidence of a strong connection idly evolved from the earlier estimation of ‘only’ fracture
between the normal stiffness (termed ‘fracture specific orientation, (with an assumed indication of major hori-
stiffness’ by the author), and the permeability, as recorded zontal stress), to fracture intensity, and the attempted
by flow rate per unit head, during ‘sector-to-sector’ flow prediction of fluid saturation and permeability
across the otherwise sealed joint perimeters. Figure 15.9 anisotropy. To make this advance, the sensitivity of frac-
shows the strong relationship between flow rate per unit ture compliance to fluids has to be understood. There
head (proportional to permeability), and the joint normal are several models addressing this important question.
stiffness, which changes due to the effect of normal stress, The authors Liu et al., 2000, investigated various
as illustrated in Figure 15.8. possible representations of fractures, for use in dynamic
The data shows that joints that support less flow modelling of fractured reservoirs. These consisted of an
obviously have higher stiffness, due also to the effect of in-plane distribution of small cracks, an in-plane distribu-
normal stress. The author suggested that the joint stiff- tion of contacts, and a thin layer with material infill. They
ness was inversely related to the cube root of the flow, were able to derive specific expressions for the fracture
and that fluid flow and the seismic response of a joint compliances ZN and ZT . Their different models indicated
could therefore be inter-related through the normal strong sensitivity of the ratio ZN/ZT to the bulk modulus
stiffness of the joint in question. of the fracture infill material, with the most rapid change
Pyrak-Nolte also found that the joints that attenu- in the compliance ratio (and values closer to 1.0) occurring
ated seismic waves most at a given stress level, also sup- when the infill bulk modulus approached zero, such as for
ported more flow, which is logical. This implies an gas-filled fractures. However with realistically small frac-
implicit link between the permeability and the rock ture aspect ratios (i.e.  0.001), much lower values than
quality Q-value and more specifically between the per- 1 were suggested, as shown in Figure 15.10.
meability and the modulus of deformation, when not Liu et al., 2000, and Liu and Li, 2001 presented equa-
reduced by clay, since Qp and the pseudo-static modulus tions for the compliance ratio (ZN/ZT), which incorp-
of deformation M (or Emass), expressed in GPa, were orated the aspect ratio of cracks, and the ratio of bulk
seen to display some remarkable, empirical similarities moduli for fluid and matrix. One of their equations,
in Chapters 10 and 13. based on earlier work by Schoenberg and Hudson,
426 Rock quality, seismic velocity, attenuation and anisotropy

(a)

Figure 15.10 Liu et al., 2000 model of in-plane small cracks, for
estimation of the ratio ZN/ZT versus the normalized
bulk modulus ( f) of the fracture infill, via compon-
ents of Hudson’s models. The matrix was a so-called
Poisson’s solid with equal Lamé constants (bulk and
shear moduli 
).

suggested that ZN/ZT  1 if the fractures were dry,


and ZN/ZT  0 if the fractures were filled with liquid.
Their interesting theory, which however is not directly
supported by the dynamic rock mechanics experience
with rock joints reviewed in the previous section, is also
illustrated in Figure 15.11a. Their equation, on which
the figure is based, is reproduced here for reference:

1 (b)
ZN 7 5  a c   k f 
 1     (15.4) Figure 15.11 (a) A fracture compliance based prediction of fluid
ZT 8 2  c c   k 
 type (brine, heavy oil or gas) based on fracture aspect
ratios and on the ratio of fluid and matrix bulk mod-
uli. Liu and Li, 2001. (b) Laboratory evidence for the
where ac and cc are the ‘long and short axes of the ele-
importance of fluid (dry or saturated) on the ratio
mental cracks in the fracture planes’, which they also ZN/ZT as a function of stress, from Pyrak-Nolte et al.,
refer to as the aspect ratios, and kf and k are the bulk 1990.
moduli of the fluid and matrix.
Figure 15.11a suggests, as does the above equation,
that ZN and ZT are zero (i.e. stiffnesses Kn and Ks are compaction due to pore collapse, fractures would truly
infinite, or very high) when a fracture is ‘closed’. Of be of infinite stiffness, since effectively absorbed in the
course if the fracture is no longer present this is logical, matrix collapse. Based on the different bulk moduli of
but not if stresses are merely high and contact areas are gas, brine and heavy oil, Liu and Li, 2001 allocated dif-
high, but significantly less than 100%. Such might be ferent ZN versus ZT ‘regions’ in Figure 15.11a.
typical for reservoir fractures in weaker materials. As partial support for the predicted rends, Liu and Li,
Perhaps for the highest-porosity chalk, showing general 2001, cite the dynamic stiffness data from Pyrak-Nolte
Shear wave splitting in fractured reservoirs and resulting from earthquakes 427

et al., 1990, and invert the ratios (see Tables 15.2 and and Lubbe, 2004, the following two equations each
15.3) to give the ZN/ZT versus normal stress trends had a small error in the first and second terms in brack-
reproduced in Figure 15.11b. They note the increased ets. In place of the 2000 versions with (4/ )  (/%a)
ratio of ZN/ZT from the saturated to the dry state. and (8/ )  (/%a), the equations are as follows:
Converting this figure to rock mechanics termin-
ology, the dynamic stiffness ratios Kn/Ks are predicted 2  
2 (r w ) 2 
1
 4
    
 

to vary from about 4 to 45 with increasing stress (when Bn    1  2  1 
rw 
 a       

saturated), and from about 2 to 4 with increasing stress 
 

(when dry). +  ( 4/3)
 (15.5)
The authors also referred to the Lucite-plate samples 

tested by Hsu and Schoenberg, in which honey-
saturated (sic) tests showed even greater sensitivity to  2
1    
the dry and saturated states, and surprisingly close mag- 2  w 2 
1
 8
    
1  2 ( r ) 

nitudes of ZN and ZT (ZN/ZT close to 1.0) when their Bt  r w    
 a   2 
2    
samples were dry, as reviewed earlier. As we have seen,  3  
 
 


the ‘intense fracturing’ (more like thin inter-bed stud-   
2 (15.6)
ies) created by the Hsu and Schoenberg, 1993 tests on (
)

either 60 or 200 (!) 0.7 mm thick, roughened (10
m) 
Lucite plates, created an extreme crack density (almost
without its equal) of 7.5 to 25, up to three orders of Here, ,  are the P-wave, S-wave velocities,
is the
magnitude greater than in situ microcracking or frac- rigidity of the rock, rw is the proportion of the fault sur-
turing. Possibly this factor is partly to blame for the face area that is estimated to consist of welded contact
unrealistic ZN  ZT result, that has permeated into (assumed 0.2), a is the mean radius of the contact
several geophysics publications. areas,
and +are the rigidity and bulk modulus of the
fault fill, and  is the mean aperture of the fault.
For purposes of comparison, one may first invert
15.6.1 In situ compliances in a these presumed micro-deformation magnitudes of geo-
fault zone inferred from physics, to the much more easily understood ‘physical’
seismic Q macro-deformation units of rock mechanics, namely
shear and normal stiffnesses of Kn  20,000 MPa/mm
The low value of seismic Q for a fault zone (or possibly (and 20 MPa/
m), and Ks  1 MPa/mm. These by
several faults) encountered in a well in the North Sea chance, or due to more inter-related micro-and-macro
(Qseis averaging 45), described by Worthington and physical processes at larger scale and at lower frequency,
Hudson, 2000, represented an abrupt increase in attenu- are similar to the values one may need to use on different
ation, relative to the Triassic and Lower Jurassic age occasions in rock mechanics pseudo-static modelling.
sandstones, siltstones and claystones that were predom- Heavily stressed rock joints, perhaps equivalent to
inantly encountered in the well. the ‘welded’ portions of the modelled fault or faults,
Worthington and Hudson modelled the effects of a have normal stiffnesses, as we shall see in Chapter 16,
down-going P-wave between 1000 and 2000 m depth, that are of this high magnitude, while clay-filled dis-
by assuming that one or up to several faults, intersected continuities will tend to have much lower values for
the transmission path. They used a theoretical compli- obvious reasons of low frictional strength. The low val-
ance model developed by Hudson et al., 1997, to demon- ues of shear stiffness that are implied by the authors’
strate the need for a very large but (to rock mechanics modelling, are in fact very reasonable values for blocks
thinking), realistic inequality of the shear and normal within faults of large dimension, that are assumed to
compliances, suggesting the need for ZN (or Bn)  4.4  dominate in large scale deformation processes, such as
1014 m.Pa1, and ZT (or Bt)  1.1  109 m.Pa1. overburden subsidence over compacting reservoirs.
The equations used to estimate these compliances are For example, Barton et al., 1986 and Barton, 2002b,
quite complicated, and are reproduced here to empha- describe (‘static’) Ks as low as 102 MPa/mm, needed to
sise the geophysical ‘way of thinking’ regarding micro- realistically model large scale discontinua in the Ekofisk
deformation compliances. (According to Worthington reservoir subsidence, thereby obtaining a better match
428 Rock quality, seismic velocity, attenuation and anisotropy

to reality than continuum models. Such low values were of data for the pseudo-static values, it may well prove
selected from a Barton, 1982 assembly of large-scale shear useful to ‘bridge-the-gap’ between the disciplines and
stiffness data (see Chapter 16). The extremely compliant ‘quantify’ () and (). This will be attempted.
values of pseudo-static shear stiffness were due to the
assumed large effective ‘block sizes’ of the major
15.7 Shear wave splitting
deforming elements of a conservatively estimated
from earthquakes
150 km3 deforming volume of overburden. (The 300 m
thick pear-shape reservoir in jointed chalk, in area
Monitoring shear waves caused by the effects of earth-
roughly 9  14 km, has a depth of roughly 3 km).
quakes requires a certain consideration of the geometry
The approx. 1.0 MPa/mm, for the dynamic shear stiff-
of the situation. There is a so-called shear wave window,
ness derived by Worthington and Hudson for the sub-
which is the area on the earth’s surface above the earth-
North Sea fault, is equivalent to a static shear stiffness
quake, where ray-paths subtend angles of incidence less
under 20  25 MPa effective stress (1.5 km gives about
than about 35°. This derives from the requirement of
37  15  22 MPa effective stress) of roughly 10 m
angles of incidence less than  sin1(Vp/Vs). (Nuttli
effective block sizes. (See Figure 16.14 in Chapter 16).
and Whitmore, 1962.) For a Poisson’s ratio of 0.25, is
An empirical expression for the macro-deformation,
about 35°. Outside this window the shear-wave wave-
pseudo-static shear stiffness, used in rock mechanics, is
forms are severely distorted.
given in Chapter 16, and is evaluated with various input
Crampin and Peacock, 2005 mention the added dis-
data assumptions, including the effect of saturated or dry
tortion if recordings are made on irregular topography.
conditions. In principle, the simple empirical equation for
They also mention the ‘helpful’ surface effect that a low
shear stiffness, from Barton and Choubey, 1977, is based
velocity layer may effectively widen the shear-wave
on the peak shear strength equation divided by the dis-
window by refracting the upcoming shear-waves
placement to peak. In other words a constant gradient of
towards vertical incidence. The effective window may
load versus shear deformation is assumed. In the case
then be as much as (2 x) 45° to 50°.
of normal stiffness, the empirical hyperbolic equation
The polarizations of the leading, faster, split shear-
of Bandis, 1980, also given in Bandis et al., 1983, gives
waves within the shear-wave window above many
the basis for normal stiffness estimation.
recorded earthquakes are typically observed to be scat-
Both these stiffness equations are components of the
tered by 10° to 20° about a direction parallel to
Barton-Bandis constitutive model, and utilise the rele-
the direction of maximum horizontal compression.
vant joint wall roughness coefficient JRC, and the joint
(Crampin and Booth, 1985). There can be many rea-
wall strength JCS, as measured or estimated input data.
sons for this scatter besides general heterogeneity with a
In the case of shear stiffness, the residual friction angle
‘mixed’ geology along the path length. Furthermore,
r is also needed. In the case of faults, we may consider
and unfortunately, the location of the main source of
JRC  0 (zero effective roughness), and r may be as
shear wave splitting is not known, which is why in-hole
low as 10° to 20°, depending on the mineralogy of the
instrumentation at up to several kilometres depth is so
clay or possible ‘shale-smear’. (For a comprehensive
attractive (see Chapter 10), to remove the attenuating
review of the sometimes very low in situ pseudo-static
near-surface layers, which perhaps could also be the
shear strengths of clay-filled discontinuities, often tested
source of somewhat differently oriented structure
at large scale at dam sites, but at low to moderate stress
and/or major principal stress.
levels, see Barton, 1973b).
A return to the important themes of dynamic compli-
ance and pseudo-static stiffness will be found in 15.7.1 Shear-wave splitting in the
Chapter 16, to better explore the similarities and dis- New Madrid seismic zone
similarities, as the case may be. At this stage of compari-
son one can assume that the pseudo-static values of Rowlands et al., 1993, using a network of more than
stiffness are lower (in the normal direction), and much thirty three-component digitally recording seismom-
lower (in the shear direction) than the equivalent eters in the New Madrid seismic zone in the central
dynamic values of stiffness. In view of the extreme lack USA, recorded shear-wave splitting, which they attributed
of present data for the dynamic values, and the wealth to EDA cracks. A compressed summary of the seismic
Shear wave splitting in fractured reservoirs and resulting from earthquakes 429

zone, an example of shear-wave splitting, and of the split ML  4 San Andreas fault earthquake; time delays at
shear-wave polarizations, together with the velocity Station MM (7 ms/km) were roughly twice those at
structure through the local crust is given in Figure 15.12. station VC (4 ms/km). The authors concluded that
The majority of the seismic events were between 3 and the relatively greater shear wave splitting observed at
15 km depth, mostly in the Precambrian basement. The station MM suggested that the fluid-filled fractures
shear waves propagating upwards encountered three within the fault zone were more extensive than in the
high-to-low velocity interfaces and one low-to-high surrounding crust, which is logical.
interface. The aligned polarizations lay in the ENE- The fault parallel polarization of the leading split
WSW direction (average N62°E), reportedly coincident shear wave at station MM was taken by the authors to
with the direction of the regional maximum horizontal indicate that either the stress was highly irregular in the
stress documented by Zoback and his colleagues. immediate vicinity of the fault, or that the fault-related
This good correlation suggested to the authors that the fractures were aligned by fault shearing rather than by
anisotropy causing the shear-wave splitting was likely to the regional principal stress.
be caused by the presence of fluid-filled, stress-aligned ver- The Liu et al., 1993, analyses were performed with
tical extensive dilation anisotropy (EDA) cracks. The time co-authors’ data, also recorded between January 1989
delays (120 to 180 ms) and associated shear wave travel and July 1990. All the earthquakes during which shear
times (3.8 to 4.6 s) suggested a shear wave anisotropy over waves were clearly visible above the P-wave coda,
the whole path of 3 to 4%, consistent with a Crampin, showed evidence of shear-wave splitting. In a few cases,
1993a, review of these ubiquitous phenomena. only one of the anisotropic shear-wave polarizations
Rowlands et al., 1993, found evidence that the was recorded, due to attenuation of the slower compo-
anisotropy causing the polarization alignments may have nent. The dominant frequency was 20 Hz, yielding
been present only in the top 5 km, rather than evenly elliptical motion. Various different polarization align-
distributed. They also reported that the basal Palaeozoic ments of some stations (Figure 15.13b) were thought to
deposits (Vp  4.83 km/s, see table in Figure 15.12) be possible due to local topography, or due to changes
were considered to be fractured, while the thick shale for- in angle of incidence due to refraction across hard rock-
mation might have introduced a thin-layer anisotropy. to-sediment interfaces.
A marginal decrease in the time delay following the
Ml  4 earthquake of May 1989, followed by an irreg-
15.7.2 Shear-wave splitting at ular increase was noted from stations MM and VC.
Parkfield seismic Similar decreases in time delay at the time of other larger
monitoring array earthquakes (M  3.5, M  6.0) were cited. The aver-
age time delay at MM on the fault zone was about twice
Liu et al., 1993 also confirmed shear-wave splitting at a as large as those at station VC 5 km away, on the south-
majority of the recording stations at the Parkfield bore- west block of the SAF. A greater density of microcracks
hole seismic network along the San Andreas fault zone, and fractures in the fault zone was cited, and change of
obtained during 18 months of recordings from previous stress affecting the geometry of these fluid-filled fea-
campaigns by co-authors Evans, Booth and Crampin, tures was assumed as a possible cause for the temporal
reported in 1984 and 1985, but recorded earlier than changes.
this. At three of the stations, distant 1 to 5 km from the A migration of focal depth for about 100 days after
fault, the polarizations were consistently normal or sub- the Ml  4 event was first cited as a possible cause of
normal to the fault strike and parallel to the direction of the decrease in time delay. However, for earthquakes
the regional maximum horizontal stress. with focal depths above 7 km depth, there was a pro-
At station MM, immediately adjacent or within the nounced increase in time delays. So the conclusion was
fault zone (see Figure 15.13, reproduced from Liu et al., drawn that the anisotropy of the fault zone was concen-
1993), first motion (qS1) was polarized parallel to the trated above 7 km.
fault strike, i.e. in this case perpendicular to the regional Since the polarizations at VC, VR and ED (Figure
maximum principal stress of about N30°E. 15.13b) were approximately perpendicular to the SAF,
There was some evidence of temporal variation of the and parallel to the regional principal stress direction
shear wave time delays in connection with a magnitude (N30°E), they were considered to be consistent with
430 Rock quality, seismic velocity, attenuation and anisotropy

(a) (b)

Vp Vs

(c)

(d)

Figure 15.12 A compressed summary of Rowlands et al., 1993, measurements and analyses of shear-wave splitting at the New Madrid seis-
mic zone. (a) seismic map, (b) shear-wave polarization examples, (c) table showing velocity structure and geology, (d) a shear
wave splitting example (focal depth 9.1 km, epicentral distance 5.0 km. Traces on left are the original, 3 seconds duration.
Traces on right are the horizontal components rotated into the fast and slow split shear-wave polarizations.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 431

(a)

(b)

Figure 15.13 (a) San Andreas fault and associated shallow thrust faults, with Parkfield high resolution seismic network (HRSN).
(b) Epicentres of local small events within the shear-wave window. (Star in SE corner was an Ml  4 event at 8.3 km depth.)
Lower hemispheres show polarizations of leading split shear-waves, beneath 7 of 9 stations. Liu et al., 1993.
432 Rock quality, seismic velocity, attenuation and anisotropy

that the anisotropic structure near Cajon Pass had


orthorhombic symmetry. ‘Stress-aligned’ fluid-filled
microcracks and pores were assumed, i.e. the standard
Crampin EDA assumption, but in fact the polarization
direction of N 13° W was reportedly not consistent
with the numerous stress measurements and borehole
break-out analyses of the surroundings, and of the upper-
most 3 km of the borehole. These had suggested N 57°
E  19°. (The strike of the San Andreas Fault, 4 km
distant, is N 60° W.) In the upper 1000–1820 m at
Cajon Pass the polarizations of split shear waves had
earlier indicated a nearly consistent N 70° E  10°.
The horizontal stress from inferred focal mechanisms
around Cajon Pass was reported as N 17° W, which was
much closer to the polarization direction N 13° W of
the authors, using ‘below 2.5 km depth’ seismic data.
The authors commented that the behaviour of shear-
waves in the vicinity of fault zones is complicated, with
the leading shear-wave polarizations often exhibiting
Figure 15.14 Superimposed fault plane solutions for earthquakes
fault-parallel alignments near the fault, but alignments
close to station MM. 95% of the shear waves propa- with the regional stress field away from the immediate
gated within 15° of vertical. Nishioka and Michael, fault zone. (This would suggest respectively fault-aligned
1990, cited by Liu et al., 1993. and stress-aligned joint sets, as the possible source of
shear-wave splitting.) The fact that stress directions
inferred from shear-wave polarizations at 2.5 km depth
were different from those inferred nearer the surface did
the extensive-dilation anisotropy (EDA) concept of
not suggest to the authors that joint orientations could be
stress-aligned fluid-filled cracks.
different, but that microcrack-controlling stress direc-
Nearer the fault, where the leading shear wave is
tions were different. The authors concluded from their
polarized parallel to the fault, the indication was of
study that the San Andreas Fault was probably driven
fault parallel cracks instead of the adjacent regional stress
by deep, regional tectonic stresses.
parallelism of the other stations. The internal structure
of the fault gouge and transition zone was assumed to
be the reason for this rotation. In the immediate loca-
15.7.4 Stress-monitoring site
tion of the fault zone, fault plane solutions for 68
(SMS) anomalies from
events close to station MM (Figure 15.14) indeed show
Iceland
a different, and quite narrowly defined maximum stress
axis between N15°W and N10°E, i.e. rotated in rela-
From another seismically active region, Crampin, 2003
tion to the regional stress along N30°E.
described the preliminary establishment of a stress-moni-
toring site in Iceland, using state-of-the-art borehole
instrumentation to monitor shear-wave splitting between
15.7.3 Shear-wave splitting a controlled-source well and two receiver wells. Such sites
recorded at depth in were designed to identify the nearly negligible changes of
Cajon Pass borehole stress, which in appropriate circumstances could monitor
the build-up of stress (or crustal adjustments) before
Liu et al., 1997 also analysed 51 local earthquakes earthquakes and volcanic eruptions.
recorded at 2.5 km depth in the Cajon Pass scientific Figure 15.15 reproduces an interesting series of
borehole, to assess shear-wave splitting. Time delays observations from the SMS at Húsavik, Iceland, from
between the split shear-waves of up to 44 ms per km were 8th to 24th August, 2001, also reported in Crampin,
identified for 32 of the events. Their analyses suggested 2003. The 12-hours (per day) histograms of seismicity
Shear wave splitting in fractured reservoirs and resulting from earthquakes 433

recorded within 100 km of Húsavik show peak activity


coinciding with an anomalous 1 m water level drop at
33 m depth in a well on nearby Flatey Island, immedi-
ately above the Húsavik-Flatey Fault. There was also
correspondence with GPS displacement anomalies of
many millimetres magnitude around Húsavik, and with
various changes in prior and subsequent P-wave and
S-wave travel times and delays.
The shear-wave splitting was the most sensitive
parameter besides ground deformation, showing a 10%
variation in time delays. Significantly, Crampin, 2003
also mentioned that with the controlled source and cross-
borehole seismic represented by SMS borehole equip-
ment, one was ‘free of the 90°-flips’ that may occur in
source zones, (due to extreme pore pressures according
to Crampin, or due to extreme loading in the H max
direction, causing an extreme ‘Poisson expansion’ due to
Crampin APE-crack expansion, according to the writer:
see earlier discussion).

15.7.5 SW-Iceland, station BJA


shear wave anomalies

Figure 15.16 reproduces the key results of a four-year


study in Iceland, reported by Volti and Crampin, 2003
for seismic recording Station BJA in south-west Iceland,
during the period 1st January 1996 to 31st December,
1999. The middle and top diagrams show the variation
of time-delays with time, for raypaths in Band-1 and
Band-2, which separate observations making solid
angles of 15°45°, and 15° to the average ‘crack’
plane. The time-delays in ms per km are normalized to
a 1 km path length.
The vertical lines through the time-delay points are
error-bars, based on errors in hypocentral distance. The
irregular curves are nine-point moving averages. The
inclined lines in Band-1 (middle diagram) are least-
squares estimates starting just before the minima of the

Figure 15.15 Some coordinated observations of changes brought


about by presumed, seismically induced crustal
adjustments, recorded at the Stress-Monitoring Site
(SMS) at Húsavik from 8th to 24th August, 2001.
The 12-hours (per day) histograms of seismicity
recorded within 100 km of Húsavik (diagram f ),
show peak activity coinciding with a variety of P-
and S-wave changes (diagrams a to c), to ground
movements (d), and to water-well level changes (e).
Crampin, 2003.
434 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.16 Correlation of time delay changes with timing of seismic events in Iceland. The inclined lines are least-squares estimates of
time-delay increases, which end with each of the larger events. Volti and Crampin, 2003.

nine-point moving average, and ending when an earth- publications, implies that even hand-specimens ‘of almost
quake or large eruption takes place. all rocks’, would be pervaded by tens to hundreds of
The arrows indicate the time of these events, with millions of microcracks, since the (radius)3 term with
date and epicentral distance. The bottom diagram shows suitably small microcracks (e.g. 10 or 100
m) guar-
the magnitudes of earthquakes greater than M  2, antees completely unrealistic numbers of microcracks
within 20 km of this recording station. to generate, for example, e  0.045 (according to
It was Crampin’s opinion that the changes observed Crampin a typical maximum crack density), as needed
in shear-wave splitting, both before the earthquakes for generating the typical maximum 4.5% shear-wave
and volcanic eruptions, were due to changes in the rock anisotropy.
mass, rather than associated with changes in the imme- In fact, as we shall see, there are reported crack dens-
diate source zones. Crampin claimed that the shear- ities and shear wave anisotropies far higher than suggested
wave splitting showed ‘that almost all in-situ crack by Crampin, when measuring shear wave polarization
distributions verge on fracture-criticality’. In the opinion through jointed or fractured reservoirs. When the crack
of the writer, the Iceland data is of great interest because density calculation is applied to the often observed scale of
of the extreme sensitivity of quite different scales of several meters (in wells and deep tunnels), quite reason-
‘crack’, namely joints and natural fractures, to minute able numbers of larger fractures or joints are predicted,
changes of effective stress and/or fluid pressure. and these features with their extremely low aspect ratios
As argued previously, this claim that almost all in-situ are, according to fundamental principles of geophysics
crack distributions verge on fracture-criticality does not and rock mechanics, surely more compliant than the
appear valid, if one assesses that the greatly expanding microcracks apparently favoured by Crampin.
microcracks in the APE model (Figure 15.3) is an improb- At great depth, as stresses are more isotropic and very
able model for rock masses. The crack density definition is high, there may be no microcracks or open fractures to
ambiguous, and as applied by Crampin in numerous speak of, as they may be completely closed. It is then
Shear wave splitting in fractured reservoirs and resulting from earthquakes 435

Figure 15.17 Left-hand diagrams: Stable sliding of several millimetres on normally loaded (6 MPa) planar-but-roughened surfaces in Stripa
‘granite’ (quartz monzonite), caused a marked reduction (up to 50%) in the shear-wave amplitude. Right-hand diagrams: Stable
sliding under a normal stress of about 10 MPa, followed by stick-slip effects, showed cyclic reduction of the shear-wave ampli-
tude in abrupt steps for each slip event. The lower diagram shows an expanded view of the ‘continuous’ sliding, showing actual
(but inaudible to AE) faster-slip episodes, which also caused quite rapid reductions of shear wave amplitude. Chen et al., 1993.

that the ambiguous, undefined location of the source The authors noted that the reduced shear-wave
depth of the shear wave splitting comes into play, and amplitude was probably associated with reductions in
one can assume with some confidence that the source (specific) shear stiffness of the joint or joints undergo-
of the splitting will tend to be shallower than perhaps ing shear. As shown by Boitnott et al., 1992, and as also
desired, where conditions of anisotropy are more modelled with the ‘JRC-mobilized’ concept of the
favourable for shear-wave splitting. Barton-Bandis joint constitutive model (Barton, 1982:
see Chapter 16), there is likely to be an ‘invisible’ shear-
ing of micron-size, prior to measurable sliding events,
15.7.6 Effects of shearing on
which could be the reason for the precursory shear-
stiffness and shear
wave amplitude reduction noted by the authors. This
wave amplitude
reduction is registered before sliding is detected, and
will logically occur before any dilation is registered,
In relation to the possible use of shear waves in detect-
with the eventual mobilization of roughness in the
ing earthquake precursors, it is interesting to note Chen
‘sliding-up’ phase, when peak strength is approached.
et al., 1993 investigations of the amplitude of shear waves
under the influence of stable sliding and stick-slip.
There were strong indications of reduced shear wave
amplitude, both under stable sliding, and during stick- 15.7.7 Shear-wave splitting at
slip events, with build-up of amplitude when ‘stuck’ a geothermal field
and rapid reductions in amplitude just before and dur-
ing slip. Figure 15.17 shows some key results. There was Shear-wave splitting at one of the largest geothermal
reportedly no change in velocity during the stick-slip sites in the world, the Cerro Prieto Geothermal Field
process, which could help to explain why Vp/Vs reduc- (CPGF), was described by González and Munguía,
tions prior to some earthquakes are not necessarily reli- 2003. Data from both weak and strong-motion earth-
able precursors for all cases. quakes was used, with seismic recording at about a dozen
436 Rock quality, seismic velocity, attenuation and anisotropy

stations, in and surrounding the production zone of the 200 m recording depth (that was assumed to have caused
CPGF. Eight of their seismic stations showed that the the shear wave splitting), was actually poor, resemblance
faster shear waves were polarized in a range of directions to dynamic moduli of deformation (expressed in GPa)
from N 14°W to N 17° E. At the four remaining sta- would be a more appropriate suggested approximation.
tions the polarization trends were between N 25° E and The authors cited microcrack alignment and their
N 67° E. For the entire area N-S was the best average. response to in situ stress as the reason for the anisotropy.
Surprisingly, in view of the need for joint connectivity However, based on the scale-dependence of dispersion
at a geothermal site, the authors followed the EDA con- seen in recent dynamic poroelastic matrix-with-fractures
vention, and assumed that it was microcracking that modelling, reviewed later in this chapter, it would seem
caused most of the shear-wave splitting. So when there that the anisotropy is more likely to be caused by
was variation, they assumed that variation of stress direc- preferential sub-vertical joint set anisotropy, which gives
tion was responsible, rather than for instance variation of anisotropy at the low seismic frequencies, as opposed to
sheared conjugate joint set directions. One of the areas microcrack anisotropy that is dominant from 1 kHz.
of deviation corresponded to the epicentre of a swarm At higher frequencies, there is insufficient time in each
of seismic activity, which would also have suggested cycle of wave motion to allow significant movement of
shearing of larger structural features. fluid, giving a relatively unrelaxed state with less atten-
In the production zone of the field, the widest spread uation, as we shall see in the model of Chapman, 2003.
of orientation (up to 90° variation according to their
equal area rose diagrams), and trending to the west, was
actually registered, suggesting to the writer something 15.7.9 Shear-wave splitting under
that might resemble conjugate joint shearing, as illus- the Mid-Atlantic Ridge
trated in Figure 15.2. Good connectivity in a produc-
tion zone, satisfying a ‘constant’ N-S average maximum An interesting combination of shear-wave splitting
stress direction, can readily be provided by conjugate analysis from local earthquakes, and P-wave anisotropy
jointing, preferably with some non-planarity if high measurement with controlled sources, was used by
permeability is to be generated due to slight dilation. Barclay and Toomey, 2003, to interpret the anisotropy
See the extensive discussion of this topic in Chapter 16. at a 35° N sector of the Mid-Atlantic Ridge, under an
ocean depth mostly in excess of 2 km. The anisotropy
was attributed to a shallow distribution of vertical,
15.7.8 Shear wave splitting during fluid-filled cracks, aligned parallel to the trend of the
after-shocks of the Chi-Chi axial valley. (Figure 15.18a). The vertical cracks give rise
earthquake in Taiwan to so-called hexagonal anisotropy, with a horizontal
symmetry axis normal to the crack planes.
Before concentrating on reservoir measurements, it is use- Most of the shear wave delay was attributed to the
ful to record the following earthquake after-shock moni- shallowest 500 m (seismic layer 2A: see Chapter 11).
toring result from Taiwan. Seismic velocity and Qseis Here the shear-wave anisotropy was from 8 to 30% in
anisotropy were recorded in the case of near-surface mon- this highly fissured layer. The authors considered that
itoring of after-shocks, following the destructive Chi-Chi isolated fluid-filled cracks at depths from 500 m to 3 km
earthquake in Taiwan, in 1999. Seismograms recorded at were too tight to be detected by the P-wave portion of
a 200 m deep station gave clear indication of upgoing their survey, but may have contributed to the shear-wave
split shear-waves with fast and slow components, indicat- delays. The authors’ analyses were restricted to shear-
ing 8% velocity anisotropy below the top 200 m. waves arriving within the shear-wave window (0° to
The authors Liu et al., 2005, estimated Qseis values in sin1 Vp/Vs). The average value of Vp/Vs was given as
the 2–15 Hz frequency band, and found values of 61 to 2.9 for seismic layer 2A (the shallowest 400 m).
68 in the fast direction, and 43–52 in the slow direction. The authors reported that the time delays ranged
As commented upon many times, such values closely from 35 to 180 ms, with an average delay of 90 ms. This
resemble possible deformation moduli (M, expressed in was reportedly similar to other studies with micro-
GPa), which can be readily estimated by the joint- earthquakes (e.g. 100–300 ms in Iceland, 100–230 ms
property-based rock mass quality Q. (M  10 Qc1/3, see in Hawaii, 10–125 ms at the San Andreas Fault). The
Figure 15.33.) If the rock quality immediately below the axial valley floor was reportedly heavily fractured and
Shear wave splitting in fractured reservoirs and resulting from earthquakes 437

Figure 15.18 (a) Bathymetric contours along a Mid-Atlantic Ridge axial valley at 35°N, showing rose diagrams of the fast polarizations. Each
rose histogram has been normalized to the size of its most populated (azimuthal) bin. Solid circles denote the 65 earthquakes.
Squares denote the six OBS instruments, which were spaced 4 km apart, covering an area of approximately 18 10 km. (1830
air gun shots were recorded during the first day, followed by micro-earthquake recording for 43 days.) Fault scarp (lineations)
are also shown. (b) Examples of the three-component seismograms before and after rotation to the fast and slow directions. (c)
Selected examples of shear-wave splitting in horizontal particle motions, with sampling points every 7.81 ms. The open circles
are the origins. Each trace is 203 ms long. Selected from Barclay and Toomey, 2003.
438 Rock quality, seismic velocity, attenuation and anisotropy

fissured, and also had faults with strike in the assumed


maximum stress direction.
The focal depths of the micro-earthquakes were
limited to 3–4 km, and the authors emphasised that res-
olution of depth-dependent structure was not possible,
but they suspected shallow structure, based on evidence
from the shallow concentration of P-wave anisotropy,
which reportedly decreased from 4% at 500 m depth
to zero below 1.5 km depth. (P-wave anisotropy was
defined as 100  (Vp max  Vp min)/Vp average, whereas,
following Crampin, 1989, S-wave anisotropy was
defined as 100  (Vs max  Vs min)/Vs max).

15.8 Recent cases of shear


wave splitting in petroleum Figure 15.19 Conversion of a P-wave source (from a sub-surface
reservoirs air-gun in water), to PS (converted) S-waves, show-
ing the different reflected and transmitted angles.
The converted waves reflect at sub-surface interfaces
Detection of fracturing by long wave-length seismic
according to Snell’s Law. An S-wave always reflects
waves, where vertical wells have given a limited sam- more vertically than a P-wave because of its lower
pling of the possible vertical or sub-vertical structure, is propagation velocity. This asymmetry complicates
seemingly a constantly developing field. The economic the acquisition and processing of converted-wave
pressure for discovery of new fields, perhaps more surveys. Barkved et al., 2004.
heterogeneous than their predecessors, and with less
favourable matrix permeabilities, is focussing increased 15.8.1 Some examples of S-wave
attention on fracture definition and understanding. It and PS-wave acquisition
is also assuming growing importance as so many fields methods
are in a mature phase and need production assistance,
with fracturing a key to maximizing the final recovery. Barkved et al., 2004 discussed the significantly different
Optimal early production strategies are also pressing for demands of recording and processing S-waves (which give
seismic monitoring assistance. a fully three-dimensional wave field), and the simpler
In contrast to the above seismic studies that mostly recording, and processing of P-waves. The ‘graphics’ of
relied on earthquakes as the natural source for their conversion of P-waves generated by the typical sub-sur-
inversion to anisotropic properties, the hydrocarbon- face, or sub-sea explosive or air-gun source, to S-waves,
bearing regions are today, by their very nature, stable was previously shown in Chapter 14 (Figure 14.27). In
areas of commercial fluid accumulation. It may be pos- Figure 15.19, the different angles subtended at an
sible to use ‘natural’ sources such as AE only later in the interface by the converted, reflected S-wave, and by the
life of a compacting reservoir, such as Ekofisk or Valhall, converted, transmitted S-wave are emphasised. The par-
where for instance, shearing has been detected. ticle motion perpendicular to the wave propagation
The type of dynamic source will depend on whether direction for the case of the converted (PS) S-waves,
the reservoir is on land or off-shore. On land it may means that they are constrained to the plane of reflec-
range from explosives, to arrays of heavy trucks with tion, and in the figure, would have particle motion in
vibrators, which can provide repeated P-wave or S-wave the plane of the page.
sources. Offshore, sub-sea-surface air-guns or explosives Three-component sensors (one vertical, two horizon-
provide the P-wave source, with conversion to PS-waves tal) are required to record the fully three-dimensional
at the sea-floor interface, and subsequent processing as S-wave field, with one horizontal component aligned
S-waves. A recent application of the suction-anchor in the direction of wave propagation. When combined
principle for sea-floor generation of S-waves (SS for with a hydrophone that is sensitive to fluid pressure
emphasis compared to converted PS) is currently being changes, the (marine) acquisition system is referred to
tested in Norway (Westerdahl, 2005 priv. comm.). as four-component (4C) technology.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 439

Figure 15.21 Split shear waves S1 (or qS1) and S2 (or qS2 )
from polarization, due to stress aligned fracturing.
The three-component wave fields were generated
by far-offset (60° or 120°) P-waves. Stenin et al.,
2002.

Variable azimuth VSP studies were described by Stenin


et al., 2002, who analysed the converted PS-wave three-
component wavefields, in order to detect and character-
ize the fractured intervals in the Archangelsk region of
Western Russia. They used far-offset shot points, for
generating P-waves, subsequently converted to PS
waves as illustrated in Figure 15.21, as a supplement to
conventional VSP with variable azimuth sources. The
two P-wave sources with 120° or 60° separation are
Figure 15.20 Schematic of walk-away VSP, showing the shear shown in the figure.
wave splitting phenomenon as polarization occurs in Converted components PS1 and PS2, and qS1 and
an anisotropic zone, caused by sub-vertical fractures. qS2 polarized waves, caused by the assumed stress aligned
It also indicates the increased time delay between the fracturing are also shown. The velocity anisotropy of
fast and slow qS1 and qS2 split shear waves, prior to Vs max and Vs min caused by the oriented fracturing, and
their acquisition by three-dimensional or four-com- by higher compression in the interpreted stress-parallel
ponent sensors in the well. Slater, 1997. direction, is also indicated. They reportedly confirmed
their seismic anisotropy analyses with core data from
offset wells that could sample the fracturing.
On land the shear-wave source can be a direct S-wave Horne, 2003, emphasised the additional advantages
(termed SS-wave, for distinction from converted PS). A of walk-around VSP, (a circular path of multiple sources
useful representation of the subsequent shear-wave at fixed offset) shown in Figure 15.22a, for improved
splitting principle, when for instance performing walk- fracture definition. In this connection he pointed out
away VSP with a shear-wave source, is illustrated in that an incident P-wave, when converted to a PS wave
Figure 15.20, from Slater, 1997, whose Caucasan reser- that passes through a plane of (horizontal) symmetry,
voir analyses will be reviewed later. such as that caused by a vertical set of fractures, will
440 Rock quality, seismic velocity, attenuation and anisotropy

15.8.2 Classification of fractured


reservoirs

3D seismic surveys, using compressional waves to gen-


erate shear-wave reflections (converted PS-waves), as
above, have found a very important role in the identifi-
cation and characterization of fractured reservoirs. With
shear-wave splitting and polarization due to the pres-
ence of the (relatively) compliant fracture properties,
the detection of basic structural information such as
fracture density, strike and (due to symmetries) dip, can
be estimated, and as we shall see, some indications of
(a)
both fluid-type and permeability may also be obtained.
It is naturally believed that stress variations and orien-
tations can also be determined, though here we have
seen some variations and therefore suggested possible
geomechanics reasons (shearing) for occasional diver-
sion from this assumption. These will be detailed more
in Chapter 16. Such deviations from the major stress
direction would be caused by prior deformation along
the more conducting, larger-scale features, or by today’s
(or yesterday’s) water-flood treatments.
(b) The shear-wave splitting component from two sets of
fractures that lie on either side of the major stress direc-
Figure 15.22 (a) Walkaround VSP with (b) polarity reversal for
converted PS waves as they pass at different angles in
tion is clearly another reason for apparent deviation
relation to strike, through a plane of symmetry. from H max, if (or because) one of the sets dominates.
Horne, 2003. Presumably we should also be open to the possibility
that if the fracture or joint set stiffnesses differ, then the
effect of a basic crack density difference between the sets
could be compromised, i.e. altered to a degree.
show polarity reversal when incident on ‘either side’ of Nelson, 1985, 2001 suggested a fractured reservoir
the fracture strike. This is illustrated in Figure 15.22b. classification that considered the dual contribution of
In all the acquisition systems, the overburden is obvi- both the matrix and the fractures on the porosity and
ously involved in wave transmission, but has tended in the permeability. The total porosity was composed of
the past to be ‘ignored’, in the sense of assuming a ver- the relative amounts of matrix and fracture porosity,
tical symmetry axis, with no aligned fracturing. and the relative levels of permeability were caused by
Tjaaland et al., 2001 emphasised that the overburden the matrix and the fracturing combined. This simple
may also display anisotropy in the form of azimuthal- scheme is shown in Figure 15.23.
dependent velocity. In traditional seismic processing Type I reservoirs are heterogeneous and anisotropic,
the overburden was assumed to be isotropic and elastic. where fractures dominate in terms of both porosity and
Due to the actual anisotropy of the overburden in permeability. In Types II and III more reserves are
relation to the assumptions, errors will be introduced stored in the matrix, but fractures control in Type II,
when the seismic data are inverted to obtain reservoir and assist in Type III. In Type IV reservoirs, fractures
parameters. They conducted synthetic modelling with still cause anisotropy and can even create barriers, and
realistic overburden anisotropy to compare results with they are assumed to provide no additional porosity or
the isotropic overburden assumption, and showed that permeability. Nelson 2001 suggested that unproductive
travel times, and P- and S-wave velocities were each wells are often a result of not recognising the signifi-
affected. If attenuation in the overburden was ignored cance of fractures, or joints, early in the development of a
in the analyses, the S-wave velocity could show errors of field. When wells were drilled based on the assumption
20% or more. of evenly distributed, matrix-controlled production (a
Shear wave splitting in fractured reservoirs and resulting from earthquakes 441

Figure 15.23 Fractured reservoir classification of Nelson, 2001,


based on the relative contributions of matrix and
fracture porosity and permeability, to the overall
permeability and productivity of the reservoir.

form of ‘fracture denial’), it was presumably inevitable


that the term ‘bright-spot’ would be the fore-runner to
full-blow fracture appreciation.
Laubach et al., 2000 proposed a fractal-based solution
to the remarkable fact, that despite the growing interest
in fractured reservoirs, seemingly worldwide, the wells
are always vertical and the target structures probably
vertical too (see also Ch. 14). As a consequence, cores
and well-logs often give little or no useful information
about the fractures. (A fortunate exception is the conju-
gate fracturing found in anticlinal structures, where
samples of both the oppositely-dipping joint sets can
hardly be avoided.) The authors referred to the use of Figure 15.24 (a) East Texas well data for microfracture orienta-
sidewall cores to help pin-point zones having high frac- tions, with comparison to the (symmetric) spread of
ture intensity; not by reaching out to the poorly sam- macrofractures from the same well and local forma-
pled fractures, but by using microfracture and diagenesis tion. (b) Scaling patterns for aperture and spacing of
data to infer the presence of the macro-fractures. microfractures through to macrofractures. Laubach
Figure 15.24 shows a comparison of the very consist- et al., 2000.
ent microfracture strikes (nearly EW), compared with the
broader but consistent orientations of macrofractures both
from the same well and from the same formation com- abundant preserved porosity. This meant that the
bined as one data set. One may speculate both about the cements did not fill the large fractures, but those below a
probable orientation of the maximum horizontal stress, certain characteristic size were completely filled. These
and about the possibility that the ‘symmetric’ range of were the target for side-wall cores due to their abundance.
strikes for the macrofractures means they are conjugate Intermediate size fractures (the emergent threshold) were
joints and possibly sheared features, perhaps with extra those filled-to-partly open, and by the nature of Figure
good conductivity as a result (see Chapter 16). 15.24b were less frequent. They termed the original aper-
Also shown in Figure 15.24 is the fractal (up-scaling tures kinematic apertures. Cases subsequently sealed were
similarity) trend of spacing and aperture. According to suffering from clogging by post-kinematic cement.
the authors: ‘Microfracture proxies for large fractures are Further observations on the subject of ‘open’ fracture
the surrogates that can provide complete, reliable, bed-by- orientations were also made by Laubach et al., 2002.
bed evidence of fracture attributes.’ The authors found They cited comparisons of measured stress directions
that in many rocks there was a diagenesis event contem- and orientations of open, flow-controlling fractures.
poraneous with the fracture development. The large aper- These showed that open fractures in the sub-surface were
ture fractures in their experience, had mineral bridges and not necessarily parallel to maximum compressive stress
442 Rock quality, seismic velocity, attenuation and anisotropy

( H max). Fractures perpendicular to this direction would be thinking of changes in the frequency of
could also be ‘open’ if partially filled with mineral synk- microcracks in the top tens to hundreds of meters. A
inematic, or post-kinematic cements. They emphasised common experience is also that the near-surface joint
that sealed fractures parallel to H max were numerous. It orientations do reflect the major horizontal stress direc-
was suggested, from experiences in both compressional tion, and consistency with permeability tensor princi-
and extensional provinces, with production data ranging pal directions may also be expected, as demonstrated by
from 2,400 to 6,400 m depth, that the divergence 3D hydrotomography by Quadros et al., 1999.
between H max and ‘open’ fractures demonstrably con- A convenient example of much higher crack densities
tributing to flow, was ‘from a few degrees to 90 degrees’. (e  N.a3/V) than apparently detected in the earthquake
studies of Crampin and others will be given here, to
introduce the alternative viewpoint that crack density
15.8.3 Crack density and shearing may be more appropriately applied to describe the fre-
of conjugate sets at Ekofisk quency of jointing or fractures that typically compose the
might enhance splitting fluid-bearing, tangible, visible structure making up the
fractured reservoir, as opposed to a focus on microc-
In earlier parts of this chapter concerning shear-wave racks. Following Leary et al., 1990, that a given number
splitting, the approximate rule-of-thumb relation of crack populations Ni of radii ai, will give a total crack
between crack density and shear wave anisotropy was density that is the sum of the densities ei, we can con-
quoted from the paper ‘Arguments for EDA’, by sider the crack density for producing parts of the Ekofisk
Crampin, 1993a. Crampin addressed the meaning of reservoir in the North Sea.
the typical 1% to 5% differential shear-wave anisotropy, Phillips Petroleum geologist’s core logging (H. Farrell,
reportedly measured in a wide range of rock types. He pers. comm. 1985), of conjugate steeply-dipping joint-
considered that the 1% to 5% was also equivalent to ing in the porous, highly productive sections of the
the generally limited range of effective crack densities, reservoir, indicated about 10 to 12 dominant (1 m
which he assumed were usually in the range 0.01  e  long) set no. 1 joints crossing a ‘1 m window’, with oppos-
0.05. Crampin noted that the percentage of differential itely dipping set no. 2 joints showing about 4 to 6 shorter
shear wave anisotropy was usually about e  100, for a joints (30–50 cm) in this same volume. In this very real
Vp/Vs ratio of about 1.7 (1.732 was quoted). case, with obviously the desirable fluid flowing towards
One may speculate whether these apparently lower the producing wells, we can estimate a much larger
crack densities implied by many earthquake studies, are crack density than apparently suggested by earthquake
a direct reflection of the sampling of ‘average rock’ studies (reportedly 0.015–0.045, Crampin, 1993a).
(mostly imprinted on the shear-waves closer to the sur- A mid-range (2D) representation of the above joint
face), and logically derived from low seismic Q and description could be estimated roughly as follows:
seismicity-prone provinces (see Chapter 10). In con-
trast, crack densities interpreted in fractured reservoirs 11  0.53  5  0.23
from split shear-wave data, may represent a ‘biased sam- ∑ e  e1  e 2 
1.0
 1.4
ple’, i.e. may be caused by a rock mass that is more
jointed or fractured than the norm at this depth. The rock mass depicted in 2D in both Figures 15.7 and
In connection with near-surface crack density influences, 15.25, is far from ‘intact’, but it tolerated an original effec-
it is of interest to note the common experience referred tive vertical stress of the order of v  62 
to by Crampin, 1993a for relatively large time delays 48  14 MPa, with a lower effective horizontal stress.
between split shear-waves to be set up in the top tens to During the first 20 years of production this effective ver-
hundreds of meters. The variations that can occur near tical stress had built up to about 38 MPa, due to the
the surface have been termed natural directivity or ND. 24 MPa pore pressure reduction prior to large-scale water-
To an engineering geologist, this phenomenon would flooding (which was followed by 6 m jack-up of all plat-
probably imply increased occurrence of jointing, and to forms, and final relocation of central platforms). In a sense
a rock mechanic with some geophysics interpretation the ‘fragmentation’ worries of Crampin at high crack den-
added, the larger time delays would also be expected to be sities was seemingly being demonstrated, but in fact it was
related to the reducing normal and shear stiffnesses of the the matrix pore-space that was collapsing in the highest
near-surface jointing. Neither profession, it is suggested, porosity chalks, supplemented by a shearing mechanism.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 443

highly jointed rock mass was well confined. It would be


correctly described as interlocked rather than ‘frag-
mented’, and it is clearly an exceptionally good reservoir
rock mass regarding fluid-flow, with planned produc-
tion possibly to 2050, an 80-year reward to Phillips
Petroleum, for not abandoning the North Sea, due to
the promising result of their final drillhole, when
exploratory drilling in the late 1960s.
Domal carbonate or chalk reservoirs of high porosity,
with steeply dipping, as opposed to flat-lying jointing, can
apparently be successful producers because of a remark-
able joint shearing mechanism, despite the one-
dimensional (vertical) strain constraint. Matrix shrinkage
under the large increase of effective stress, ‘makes space’ for
down-dip joint shearing. The latter helps to maintain joint
aperture due to shear-induced dilation, and apparently
may even provide a pseudo-confinement effect (increased
(a) ko: ratio of horizontal to vertical stress), which would make
the jointed reservoir somewhat stiffer (in a vertical direc-
tion), than the unjointed rock (Barton et al., 1985, 1986).
Significantly, in view of the fact that our shearing theo-
ries were not at first believed, Albright et al., 1994 men-
tion Ekofisk exhibiting ‘shear fracture microseismicity,
possibly indicating that subsidence is caused by a combi-
nation of pore collapse and shear sliding’. They state that
subsidence surpassed early model estimates based on
pore collapse, indicating other mechanisms at work. By
implication shearing was also occurring at fault scale.
(See further discussion of this shearing mechanism at
joint, or fracture scale, in Chapter 16.)
In this connection one should refer to possible doubt
concerning the use of laboratory 1D-compaction data
for modelling pore-collapse versus porosity. One should
recall that the modelling of compaction applies to some
30 km3 of reservoir chalk, about twelve orders of magnitude
(b) larger than the laboratory samples. Some level of scale
effect may be present in the matrix, as certainly found for
Figure 15.25 (a) A UDEC-BB model of a compacting, conju- the case of joint strength, though with strongly reduced
gately-jointed 1  1 m ‘element’ of the Ekofisk
scale effect at high effective stress levels. The relative
chalk, which has a depicted crack density of the order
contributions of the matrix and the joints (and faults) to
of 1.4. The modelled compaction of this idealized
high-porosity ‘element’ of the reservoir was 4.8%. (b)
the over-all compaction is therefore inevitably uncertain
Joint shearing (shown proportional to line thickness) where such large volumes are involved. The operator’s
was modifying the compaction in relation to that core-logging geologists reportedly detected slickensides
of an unjointed porous matrix. M. Christianson on conjugate joint sets, when drilling new holes during
UDEC-BB modelling: Barton et al., 1986, 1988. the 1980s for pressure maintenance, using equilibrated
sea-water injection. Slickensides had reportedly not been
When discretely modelled (as in Figure 15.25), the detected in earlier characterization of the Ekofisk field in
pore pressure reduction (in 1985) was limited to 20 MPa. the late 1960s, where production started in 1971.
A corresponding increase of the minimum horizontal The most porous chalk (n  40%) that was first mod-
effective stress, plus deformation effects, means that the elled by the NGI team in 1985, with a modulus of only
444 Rock quality, seismic velocity, attenuation and anisotropy

0.33 GPa, showed maximum shearing of 3.9 mm, with production, due to the low matrix permeability of the
an average of approximately 0.39 mm for all the joints. respective chalks, despite their high porosity.
Later modelling by Gutierrez, with a higher pore pressure Leary et al., 1990 referred to exceptional 2.44 km/s and
reduction of 24 MPa, showed up to 10 mm maximum 1.83 km/s fast and slow shear-wave velocities interpreted
joint shear. Presumably such deforming features would parallel and perpendicular to the inferred alignment of
be seismically visible in practice (i.e. a strong source of fractures, in an oil-bearing stratum in the same Austin
shear wave splitting), just as the reproduced direct shear chalk. The shear-wave velocity anisotropy reported by
tests on rough fractures (Barton, 1973) shown in Figure
15.26 are ‘visible’, due to the dilation that is caused.
Water flooding may have stimulated this shearing
mechanism due to intuitive, preferential weakening at
highly stressed joint-wall contacts, prior to water-
softening of the less permeable matrix. (Barton,
2002b). However, the Ekofisk rock mass does not ‘frag-
ment’ or suffer ‘dispersion’ of pore fluid because of high
crack density, although the weakening effect of the
water is tending to cause ‘rubble-isation’ of the matrix-
plus-joints, but this is outside the scenario envisaged by
Crampin, 1993a. The oil undoubtedly flows more eas-
ily towards the producing wells in response to the com-
paction drive and due to the originally well-developed
crack-density-determined connectivity and resulting per-
meability. Loss of matrix (and joint) strength due to
water-saturation, gives an additional production effect.
The authors Saenger and Shapiro, 2002 used an
explicit finite difference scheme to model elastic waves in
their variously ‘cracked’ models. Interestingly, consider-
ing the above independently estimated crack density
e  1.4 for the Ekofisk conjugate jointing, derived by
the writer from a geologist’s description of the jointing, is
the fact that Saenger and Shapiro find a value of e  1.43
to represent the critical crack density. They liken this level
of crack density to a medium with ‘only finite sized pieces
of solid, and there is no continuum through which an elas-
tic wave can propagate’. To one from a different back-
ground (rock mechanics), this would almost seem like an
unintended definition of the typical near-surface jointed
rock masses through which we are frequently performing
seismic refraction, core drilling, and driving tunnels,
with frequent need for rock support.
From another chalk reservoir, in Texas, relative levels of
shear-wave anisotropy detected by VSP above the Austin
Chalk reportedly correlated with production (Crampin,
1993a), as one would certainly expect at Ekofisk. A ten-
dency for near-surface layers to reflect this same
anisotropy was also noted for the Austin Chalk. Dim-
ming of the amplitudes of the slow shear-wave, (due to
greater attenuation i.e. lower Qseis), correlated with frac- Figure 15.26 Reconstruction of the measured shear-dilation path
tured zones, which were verified by horizontal drilling. from direct shear tests of rough tension fractures. n
The detection of the fractured zones was vital for good 2 and n 6 depict stress level. Barton, 1973a.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 445

Johnston, 1986 of about 30%, would have implied a Horne et al., 1997 described each anisotropic zone
crack density of the order of 0.3, based on the Crampin using three parameters: the crack density, the aspect
‘100e’ relation, which presumably requires modifica- ratio and the crack content, and a variety of advanced
tion if there are more than one joint or fracture set con- operations which we need not be concerned with in
tributing to the crack density. The ratio Vp/Vs , and the this simplified review. The authors’ results supported
relative levels of compliance of the contributing joints earlier studies that also concluded that the observed
(not directly mentioned by Crampin due to microcrack anisotropy at the CBTF well was likely to be due to a
focus), will also alter this ‘100x’ relation. sub-vertical fracture set dipping approximately 18°
If the crack density is contributed to by two sets of (to the SE). They showed that the fracture dip could
oppositely dipping conjugate joints, then following be obtained from the shear-wave anisotropy when
Sayers, 2002b, we can expect both shear and normal using opposite-azimuth VSP. The authors’ representa-
compliance contributions (from both sets), to the slow- tion of the velocity anisotropy structures of qP, qS1
ness of the slow shear-wave, and presumably therefore a and qS2 wave forms for the case of vertical cracks (or
lesser need for an extreme (interpreted) crack density to joints) and 20° dipping cracks (or joints) is shown in
‘explain’ a high value of shear-wave anisotropy. Figure 15.28.
In this last case reviewed we saw how the use of
opposed-azimuth VSP could be used to interpret the dip
15.8.4 Links between shear wave
anisotropy and permeability

Horne et al., 1997, presented what they term ‘a global


optimization’ to address the problem of ray-tracing, to
invert the observations of shear-wave splitting from two
‘near-offset’ VSP data sets, in which two sources were
employed along diametrically opposite azimuths, about
the wellhead.
Their shear-wave splitting analyses were based on
shallow depth VSP and logging data for the Conoco
Borehole Test Facility (CBTF), where an 18-layers inter-
pretation of velocities, densities and thicknesses was in
use by researchers. This layered model was split into
five anisotropic zones that corresponded to the observed
discontinuities in the shear-wave splitting estimates, as
obtained from the diametrically opposed VSP surveys.
Figure 15.27 shows the basic elements of the velocity
density model, and the five selected zones.

Figure 15.28 A comparison of anisotropic velocity components, if


Figure 15.27 The velocity and density model for the CBTF well. vertical (left) or 20° dipping cracks (right). Lower
Five anisotropic zones were identified from the plots show time delay variations over a hemisphere
shear-wave splitting analysis. Horne et al., 1997. of propagation directions. Horne et al., 1997.
446 Rock quality, seismic velocity, attenuation and anisotropy

of the fractures. The response will be a complicated


function of aspect ratio (which includes a normal stress
component), fracture fill and/or porosity. A minimum
of three source-receiver lines are capable of detecting
fracture strike, if the particular set of fractures gives
directional velocity variations in relation to the strength
of other sets. The response is strongest when fractures are
gas filled, as we shall also see later from a case record
from Oman. In addition, a lower normal effective stress
perpendicular to the major set of fractures will give a
stronger velocity contrast. The gas to oil ratio in the
reservoir is therefore an important component of the
interpretation.
The authors pointed out that the new generation of
vertical cables, seabed seismic sensors and walk-away
(and/or 3D) vertical seismic profiling was leading to high-
resolution anisotropy estimation, specifically in the off-
shore environment. The common goal was to detect this
commercially important azimuthal anisotropy. If the
Figure 15.29 Azimuthal connection is indicated between the alignment of fracture sets could be deduced, this would
degree of seismic anisotropy (‘birefringence’) and
assist in the optimal lay-out of producing wells, and per-
the permeability. Several anisotropic domains are
haps subsequent injection wells for water-drive.
indicated. (Horne and MacBeth, 1996, as repro-
duced in MacBeth and Li, 1999.)
Azimuthal anisotropy detection techniques can help
resolve both the large fault-scale structures, and the
expected inter-block scale fracture (or joint) set struc-
of fracture sets or joint sets. The more usual application tures. The advantage of this ability to interpret struc-
of shear wave splitting is to determine the azimuth. ture more fully, at several scales, can be envisaged in the
Figure 15.29, again from the Conoco test site, shows complex stratigraphy shown in Figure 15.30.
the Horne and MacBeth, 1996, linkage between seismic MacBeth and Li expressed the opinion that azimuthal
anisotropy (quite often termed birefringence) and the anisotropy determination at fracture set scale, could fill
recorded permeability. We will recall that there were five the gaps between the fracture characterization from core,
domains (or zones) of anisotropic wave-splitting at this from borehole logs, from outcrop analogues, and that
site. MacBeth and Li, 1999 also emphasised that the rela- inferred from 3D seismic above complexly faulted and
tionship between the stress field and a fracture model fractured reservoirs. The conceptual scale-dependency
may require investigation from case to case. It appears of fault throw magnitudes, after Yielding et al., 1992
that the hydraulically connected fractures may be those and MacBeth and Li, 1999, is sketched in Figure 15.31.
most readily detected, which implies that the link between Interesting from a rock mechanics viewpoint was the
permeability and anisotropy is being detected also. suggestion by MacBeth and Li that ‘sub-vertical wave
However, the generally assumed influence of a nearly propagation through vertical fractures could provide
parallel major horizontal stress, and therefore min- direct knowledge of the conditions influencing tangential
imum normal stress must be assessed carefully, as we movement across the fracture faces’. They suggested that
shall see in Chapter 16. It may follow present conven- this could lead to discrimination between the solid or
tion in shallow wells, while a (possibly conjugate) shear- liquid content of the fractures, evaluation of porosity,
ing mechanism may be needed at depth, unless rock and perhaps geometric aspects of the surface topogra-
strength, or partial mineralization ‘bridging’ are sufficient phy. One may speculate if the numerically modelled
to keep fractures ‘open’, despite a significant effective joint shearing shown earlier for an element of the
normal stress. Ekofisk jointing (Figure 15.25) could be seismically
For vertical joints or fractures, the most pronounced detected, likewise the rock-to-rock (R) and fluid-
anisotropy is registered when the source-receiver propa- bearing open (O) parts of shearing joints depicted in
gation paths are parallel and perpendicular to the strike Figure 15.2b.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 447

Figure 15.30 A seismic interpretation that shows complex stratigraphic and structural relationships (London-Brabant Massif, North Sea).
Seismic azimuthal anisotropy analysis offers potential for resolving major fracture set azimuths within the faulted blocks.
MacBeth and Li, 1999.

The authors found that the shear wave data could


be interpreted as showing both velocity anisotropy
(azimuthal variation in travel times), and attenuation
anisotropy (azimuthal variation of amplitudes). The com-
mon central azimuth N70°E, which was also the strike
of the dominant, steeply dipping macro- and micro-
fractures, happened also to be the direction of maxi-
mum horizontal stress at shallow depth in the area.
(Figure 15.32a, b and c). At this shallow near-surface
location there was clearly excellent stress-fracture
alignment, as expected.
The nearby surface exposures suggested two approxi-
mately perpendicular fracture sets at the CBTF. Liu
et al., 1993 managed to infer the multiple fracture sets
using shear-wave polarization, with sufficient azimuthal
coverage, in this case 160° of data. Their equal area
polar projections show two maxima in approximately
orthogonal directions, which almost exactly parallel the
Figure 15.31 MacBeth and Li, 1999 suggested that seismic strikes of the two fracture sets. Figure 15.32a shows the
anisotropy analysis could fill a gap between the frac- outcrop jointing and Figure 15.32b the velocity struc-
ture characterization from core and that from bore- ture. The target for these analyses was a fractured lime-
hole logs and outcrop analogues, and that inferred stone from 16 to 29 m depth, sealed above and below
from 3D seismic. From Yielding et al., 1992. by impermeable shales. Its velocity of about 4 km/s
does indeed suggest well fractured rock (but a small
velocity-depth gradient across this layer is perhaps
15.8.5 Polarization-stress likely).
alignment from shallow Reference to Figure 15.33, which reproduces key rock
shear-wave splitting mass quality Qc-Vp-depth charts, suggest a rock quality
in the range 2  Qc  4, when a porosity adjustment
Liu et al., 1993 had previously addressed shear-wave of presumably some few per cent is made. This typical
sources from the shallow (50 m deep) multi-azimuth, jointed rock quality, could be ‘generated’ by a logical
reversed VSP at the Conoco Borehole Test Facility, where combination of e.g. RQD  45–90, Jn  9 (3 sets of
earlier analyses had demonstrated that fracture parame- joints; two sub-vertical, plus bedding), Jr  2 (smooth
ters could be estimated from cross-hole, reversed VSP, but undulating joints), Ja  2 (slight weathering),
and borehole data, as we saw in Chapter 14. Jw  0.66 (wet, some water flow), SRF  2.5 (near
448 Rock quality, seismic velocity, attenuation and anisotropy

(a) (b)

(c)

Figure 15.32 (a) Outcrop jointing, (b) assumed velocity structure, (c) polarization of shear-wave arrivals. Arrowheads show actual dominant
fracture strike. Liu et al., 1993.

surface, some looseness in relation to greater depth). If the uniaxial strength of the limestone was e.g.
(See Appendix A for full Q-system rating tables.) 100 MPa, then Qc (Q  c/100) would be 1 to 3,
which is mutually consistent with the measured veloc-
ity of about 4 km/s.
45  90 2 0.66 A simple approach is to check along the ‘4 km/s diag-
Q     1.3  2.6
9 2 2.5 onal’ in the simplified near-surface (nominal 25 m
Shear wave splitting in fractured reservoirs and resulting from earthquakes 449

(a)

(b)

(c)

Figure 15.33 (a) Rock mass quality Qc-Vp-depth gradients, (b) Qc-Vp-depth-porosity-modulus, (c) simplified Q-Vp- c chart (nominal
25 m depth). Based on Barton, 1995 and Barton, 2002a. Note: support pressure in b) refers to tunnels.
450 Rock quality, seismic velocity, attenuation and anisotropy

depth) model shown in Figure 15.33c. Note that in the between seismic anisotropy and productivity. He used
area of Q  2 to 3, the implied uniaxial compressive VSP walk-away data that was shot at two North Caucasus
strength is around 100 MPa, which appears realistic for a oil fields. Due to the usual low permeability of clay rocks,
limestone. If in reality the rock was closer to say chalk, hydrocarbon reservoirs in argillaceous rocks are not
then higher matrix porosity, lower uniaxial strength, and commercially viable unless they are strongly fractured.
slightly less jointing than assumed above (giving a higher Prior to the commercial use of shear-wave splitting,
rock quality Q-value, could still be invoked to derive a primarily since the early 1990s, argillaceous reservoirs
suitable P-wave velocity of about 4 km/s. were reportedly found almost accidentally, while drilling
Great azimuthal consistency is shown by this Liu et al., towards prospects in other rock units.
1993 data set, due perhaps to the shallow depth, which Slater, 1997 investigated the azimuthal anisotropy to
allows the ‘open’ joints or fractures to indeed follow the check if the strong azimuthal variation in productivity
conventional concept with likely parallelism with the was caused by variations in fracture intensity. In one of
H max direction. As pointed out earlier, and analysed in the reservoirs, walk-away VSP at two of the wells indi-
more detail in Chapter 16, such parallelism may be com- cated velocities increasing as the direction of propaga-
promised at reservoir depths in less competent rock than tion moved away from the vertical direction. The
limestone, due to effective normal stress induced joint reservoirs were in layered clays.
‘closure’. A (conjugate) shearing mechanism might then In Figure 15.20, introducing the topic of shear wave
be required for sufficient ‘openness’ of the conducting splitting, a schematic of the walk-away VSP was shown,
joints. Alternatively, suitable amounts of mineral filling from Slater 1997. Shear wave splitting and polarization
for ‘bridging’ might be needed, neither too much ( seal- was occurring in the near-surface anisotropic zone, caused
ing), nor too little ( stress-closure). by sub-vertical fractures. Also indicated is the increased
A shear wave polarization component (from e.g. two time delay between the fast and slow qS1 and qS2 split
similar conjugate joint sets) that was still roughly consis- shear waves. Later in this chapter, we will see many more
tent with the h max direction, might be assumed from examples of these phenomena, and also see models capa-
much deeper sets of perpendicular break-out analysis, from ble of simulating the squirt-flow causes of dispersion or
caliper logs. But here there would be the possibility of a frequency dependence at different fracture-size scales.
lot of ‘noise’ in the data, possibly from joint-induced Slater, 1997, examined the typical transverse isotropy
break-out at other angles, more consistent with the possi- of the lower strata in these clay reservoirs, which display
ble conjugate, partly sheared (minor-faulted) jointing. In isotropy about a vertical axis of symmetry as shown in
addition to this possible ‘anomaly’ in relation to conven- Figure 15.34. The VSP analyses demonstrated strong
tional thinking, there is also the possible ‘rotation effect’ increases in velocity when the direction of propagation
caused by the joint shearing per se, giving different orien- moved away from the vertical direction.
tation for the fluid-bearing (O) and stress-bearing (R) The very low velocities calculated from VSP at wells
parts of a non-planar, shearing joint (Figure 15.2b). 85 and 87 at one of the Caucasan oil fields are shown in
The above aspects will be addressed in more detail in Figure 15.35. The layered sandstones, limestones and
Chapter 16, where important supporting case records will clays, and the Maicop Clay, do not follow the rock qual-
be found for the ‘shearing anomaly’. It is possible that par- ity Qc-depth-Vp trends shown elsewhere in this book,
allelism of one conducting set of joints to h max is in fact except in the top few hundred metres where azimuthal
the anomaly, when no longer close to the surface, in view anisotropy indicates jointing that would tend to show
of the stress-induced joint closure phenomenon, unless closure with depth as in the Qc-Vp-depth model of
the reservoir rock is rather strong, like limestone, or a Figure 15.33. Perhaps below this anisotropic zone, the
hard sandstone but not chalk, or poorly cemented sands, change to clay, and the reported high pore pressures
which would have sufficient matrix flow to make shear cause under-compaction, despite depth increase. A lack
wave splitting a less relevant mechanism. of fit with the writer’s Q-Vp-depth-soft-porosity model
would then be logical, unless a correction for over-
pressure was made, and an ‘apparent depth’ estimated.
15.8.6 Shear-wave splitting in An important finding from careful analysis of much
argillaceous rocks earlier multi-offset VSP in the Paris sedimentary basin,
indicated non-parallel shear wave polarizations at dif-
In a doctorate study at the University of Edinburgh, ferent azimuths and offsets. Polarization by sub-parallel
Slater, 1997, also investigated if there was a relationship cracks or fractures was not sufficient to explain these
Shear wave splitting in fractured reservoirs and resulting from earthquakes 451

reservoir, was reported by Roche et al., 1997. The


Vacuum field is a fractured dolomite reservoir in New
Mexico, and the survey was a repeated 3C (converted
wave) survey, i.e. 3C  3C, or 9C survey. The argument
was made that because the dolomite was quite hard it
should be a poor candidate for conventional time-lapse
studies, (i.e. its joints probably had high JCS or joint
wall compression strengths, and perhaps high joint
roughness JRC: see Chapter 16 for the influence of JCS
and JRC on normal stiffnesses, the approximate inverse
of dynamic compliance). The joint parameters used for
pseudo-static estimation of both normal and shear stiff-
ness (i.e. JRC and JCS), are relevant here due to the
pseudo-static nature of the planned CO2 flooding, and
its effect on the increased pore pressure and therefore
reduced effective stresses.
Before the CO2 injection, the distribution of the shear-
wave splitting anisotropy parameter () varied between
/8%, but averaged about 4%. Note: () as given by
Thomsen, 1986 is defined from the elastic stiffness
matrix as (c44  c55)/2c55, where qS1  (c44/), and
qS2  (c55/). After the injection of CO2, the pore
pressure at the injector well increased from 10.6 to
17.0 MPa, giving a maximum Pp of 6.4 MPa. This
was sufficient to give a P-wave velocity decrease of only
4%. The shear-wave anisotropy parameter () on the
other hand changed by up to 14% (from 4% to
10%). These changes were successfully modelled by
Angerer et al., 2001. The polarized fast and slow shear-
Figure 15.34 Examples of transversely isotropic layering in argilla-
ceous/clay reservoirs, where VSP analyses demon-
wave velocities did not rotate during these time-lapse
strated strong increases in velocity when the changes, they apparently exchanged places.
direction of propagation moved away from the verti- According to a review of Winterstein et al., 2001,
cal direction. See Slater et al., 1993 and Slater, 1997, including 23 wells in six locations in NW USA, and
for details of the ‘anisotropic cuspidal phases’ discov- not specifically related to fractured reservoirs, shear-
ered in these studies. wave splitting anisotropy as defined above, was com-
monly in the range 0    21%. The general trend of
data, and Bush, 1990, demonstrated that the anom- the anisotropy data suggested a certain consistency in
alous behaviour could be the result of a combination of one layer or domain, with an abrupt transition to dif-
matrix anisotropy due to layering (termed ‘azimuthal ferent values. One could argue that such could be caused
isotropy’, with vertical axis of symmetry, as in Figure by sedimentary beds of different geological ages with
15.34), and crack anisotropy. A paper by Bush and differently oriented joint sets, and also by local stress
Crampin, 1991, showed the consistency of this combin- rotation near fault zones.
ation of mechanisms at five of the offsets. Some time-lapse surveys using P-waves, with its rela-
tively more simply processed information, were described
in Chapter 14. At the Ekofisk reservoir, the gas-cloud was
15.8.7 Time-lapse application seen in Figure 14.29 to obscure P-waves from a large por-
of shear-wave splitting tion of the central, and most porous part of the reservoir.
over reservoirs In the context of the newer S-wave technology, Barkved
et al., 2004 referred to the world’s first time-lapse, marine,
Reportedly the first time-lapse survey designed to measure multicomponent (3D/4C) survey, as that performed in
possible changes in shear-wave splitting above a petroleum September 2002 at the Ekofisk jointed-chalk reservoir
452 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.35 Very low shear wave and P-wave velocities at two wells, as interpreted from VSP. The geological description of the 0.75 km of
overburden is also seen. There appears to be over-pressure in the Maikop Clays. Slater, 1997. Alternatively, the reduced sensi-
tivity to effective stress may be due to low pore compressibility for the clays. Holt et al. 2005.

in the North Sea. This baseline was subsequently com- According to Barkved et al., 2004 (with ‘Ekofisk-
pared with a monitoring survey acquired in December author’ Van Dok as one of the co-authors), the reasons
2003. In each case, seabed cables were used to acquire data for the small changes detected by the S-waves ‘had yet to
with a wide range of azimuths. (Van Dok et al., 2004). be understood’. The small-scale joint-shearing mechanism
With ‘only’ about 3 108 m3 of oil out of a total of identified in distinct element (UDEC-BB) studies for
about 1.1 109 m3 produced by 2003, and production the Norwegian Petroleum Directorate (Barton et al.,
expected to 2050, it is clear that the ‘belated’ application 1985, 1986), that was discussed earlier, was later ‘con-
of shear-wave technology still has an important role to firmed’ by slickensided conjugate joint faces, in core
play. From each survey, converted PS-waves were recovered from subsequent wells (post 1985) for water-
analysed to determine the principal directions of fast flooding and production.
and slow shear-waves. At every sensor location, According to Phillips Petroleum Co. geologists this
recorded traces were binned (collected) into 10° slickensiding had not previously been noted, and nor
azimuth sectors, and then stacked, giving 36 traces of was it noted in the older jointed cores made available to
each component at every receiver location. Even this NGI for laboratory direct shear and coupled shear-
limited time-lapse of 15 months indicated some small flow-temperature (CSFT) tests. The shear mechanism
changes in the direction of the fast shear wave, and in the may seem surprising in view of the 1D-strain (‘roller-
difference between fast and slow shear velocities. The boundaries’) boundary condition, since a 9  14 km
differences were not consistent across the field. reservoir of 300 m thickness can hardly expand laterally
Shear wave splitting in fractured reservoirs and resulting from earthquakes 453

during compaction: this occurs more in the stretching,


subsiding, overburden as ‘seen’ by shallow shear-wave
splitting at Valhall, to be reviewed shortly.
A possible explanation for the small changes of polar-
ization direction and of shear-wave anisotropy at Ekofisk,
can perhaps be found in the conjugate (or single) shear
mechanism that was illustrated in Figure 15.2b. This
mechanism was also invoked earlier in this chapter, as a
possible explanation of a larger polarization rotation at
the Cornwall hot dry rock geothermal project. With
Figure 15.36 Shear-wave splitting and polarization results for the
potential ‘opposite-rotation’ of fluid lenses and rock-to-
shallow overburden above the compacting Valhall
rock contact areas (Barton, 2005), there could be subtle
reservoir. Lines show the qS1 direction, with their
domination of effects from the primary relative to the length corresponding to the qS1  qS2 time delay or
secondary conjugate joint set, and if the ‘O-R’ mech- ‘lag’. The ‘rotation’ presumably corresponds to the
anism can be detected by shear waves, and if the strike relative ‘visibility’ of sub-vertical (bedding-limited?)
of the two conjugate sets is not equally oriented, then a joints caused by ‘stretch’ in all directions. Olofsson
small rotation could be explained. and Kommedal, 2002, also Gaiser and Van Dok,
Variation about the field, with the ‘radially’ trending 2003 and Barkved et al., 2004.
jointing and rotating principal stress (Figure 14.31)
would easily explain variation of such trends. Others
might quote EDA-(micro)-cracks and stress rotation as time delay of the slow shear wave in relation to the fast
the possible cause. More subtle mechanisms may be at wave in the reservoir, where both the polarization direc-
work, and additional complications in the neighbour- tions, and the time delay correlated with the geological
hood of fault zones are almost inevitable. model of fracturing. In this particular paper they pre-
At the Valhall Field, quite close to Ekofisk, BP installed sented the first results of shear-wave splitting in the shal-
a permanent seabed cable array, covering 45 km2 area, to low overburden, indicating a remarkable, and very
monitor changes using regularly repeated 3D multicom- convincing match to the assumed ‘stretch’ of sub-verti-
ponent seismic surveys, to help determine the best reser- cal jointing caused by subsidence. (Mention of these
voir drainage strategy. The plans for this installation were mechanisms was made in Chapter 14, for the case of
alluded to in Chapter 14. As is well known, and as will Ekofisk overburden velocity reduction). Figure 15.36
also be indirectly demonstrated in Chapter 16 using shows the result of their shallow overburden shear-wave
Barton-Bandis joint closure-permeability modelling, it is polarization, with lines showing the qS1 direction, with
all too easy to produce too fast thereby prejudicing the their length corresponding to the qS2 time delay or ‘lag’.
permeability of the rock joints and fractures, close to pro- Barkved et al., 2004 also commented on the above
ducing wells, where pore pressure reduction may cause near-surface Valhall result, and stated the following: ‘The
‘too high’ effective normal stress on the producing frac- actual mechanism causing the shallow shear-wave splitting
tures. A slower production helps to maintain the vital per- is not known. Azimuthal anisotropy is usually associ-
meable routes through the reservoir, especially where ated with fracturing, stress or lithology. In this case the
matrix porosity is superior to its permeability. amount of anisotropy is small at the centre of the field,
In such modelling, the joint-roughness-dependent where the subsidence is largest, but the anisotropy is
conversion between mechanical aperture (following pore large on the flanks and small again farther from the
pressure-induced effective stress increase) and the con- centre. This strongly points to shear-wave splitting
ducting aperture, needs to be differentiated. Fortu- being sensitive to changes in stress or strain.’ By model-
nately, the loss of mechanical aperture (E) occurs more ling changes of triaxial stress in (continuum) layers
rapidly than the loss of the smaller hydraulic aperture above similar compacting reservoirs, Herwanger and
(e), as shown by Barton et al., 1985, Barton and Quadros, Horne, 2005 produced similar, but quite circular mod-
1997. (See Chapter 16 for review of E  e data sets). els of shear-wave polarization.
A very interesting application of shear-wave splitting The referred authors have not apparently focussed on
and polarization was described by Olofsson and intra-bed jointing as the likely source of the partial
Kommedal, 2002. They referred first to the significant ‘squareness’ of some of the strongest anisotropy (i.e. the
454 Rock quality, seismic velocity, attenuation and anisotropy

‘NNE-trending’ and longest lines). The depth giving


this possible dominant ‘imprint’ to the polarization and
velocity anisotropy is of course not known. Large-scale
(axisymmetric, 10 km radius) distinct element model-
ling of the Ekofisk overburden response to modelled
compaction, using numerous coarsely ‘bedded-jointed-
and-faulted (2D) UDEC models (Barton et al., 1985,
1986), showed distinct ‘joint’ opening and some ‘bed-
ding’ shear in the overburden, due to the stretch caused
by the subsidence. These effects worsened with increased
compaction profiles, the subsidence/compaction S/C
ratio exceeding 0.85 as compaction approached 10 m.
A 3D version of such discontinuum modelling (with
the 3DEC code, also developed by Cundall), would
obviously have shown similar reactions from other Figure 15.37 (a) Temporal variation in minimum % anisotropy
and b) temporal variation in differential attenuation,
perhaps perpendicular ‘joint’ sets. This ‘joint’ opening
from shear wave splitting analysis using AE recorded
occurred most strongly where bending was strongest,
in siltstone caprock, above the compacting Valhall
and least both centrally and further out beyond the jointed chalk reservoir. Carter and Kendall, 2005.
flanks. It is therefore suspected that the ‘stress or strain’
referred to by Barkved et al., 2004 could rather be
termed intra-bed joint-opening effects, since the strength results were surprising in two ways: 1) that anisotropy
and location of these phenomena are likely to match the appeared to be temporal, 2) that sometimes qS2 was
subsidence-bowl shape at Valhall. Where the polarization richer in higher frequencies than qS1.
is ‘diagonal’, (i.e. ‘NE-SW’ or ‘NW-SE’ relative to the These two unusual results are illustrated in Figures
‘N-S’ page), presumably the components to polarization 15.37 and 15.38. The authors defined a differential
(i.e. joint compliance) from both sets could be operating. attenuation, as the difference in energy loss per cycle
(See also polarization at 300 m depth, in Barkved and experienced by qS1 and qS2. This approximates the
Kristiansen, 2005.) 1/Qseis that a homogeneous constant Qseis material
would require, in order to produce the observed differ-
ence in frequency content between qS1 and qS2.
15.8.8 Temporal shear-wave One may speculate that opening of sub-vertical, bed-
splitting using AE from limited jointing, with changing joint-wall contact char-
the Valhall cap-rock acter, could be responsible for such variation over time.
According to Barkved and Kristiansen, 2005, sea floor
An unusual petroleum reservoir case record concerning subsidence at Valhall exceeds 5.4 m, approximating
temporal variation of attenuation, was described by 0.25 m/year. The larger neighbouring Ekofisk field has
Carter and Kendall, 2005. This concerned the utilisa- suffered significantly larger subsidence, at an early rate
tion of micro-seismic events generated in the siltstone of about 0.45 m per year when detected in the mid 1980s,
(above-shale?) caprock, above Norway’s Valhall jointed and about 10 m of total compaction with large num-
chalk reservoir, in rock at about 2 km depth. Subsidence, bers of sometimes repeated casing damage, by the turn
observable at the sea-floor (see previous review), caused of the century. Since the shear-wave phenomena may
this AE activity, which was recorded 300 to 500 m be relying on vertical jointing, the question would be
away, by a vertical string of six 3-component geophones, whether bed-slip could be affecting the successive open-
placed in an abandoned well near the crest of the anti- ing and closing of bed-limited jointing.
clinal structure. In the case of the Wilmington field, under Long Beach,
Over a period of 56 days, continuous recording gave California, a significant seismic event occurred, with a
572 events, 324 of which were located reliably (Dyer 20 cm presumed bed-slip, at one stage during more
et al., 1999). Carter and Kendall performed shear-wave than 10 m of subsidence. Dussault, 2001 described the
splitting analysis, comparing relative frequency content sand/shale interfaces above the Ekofisk reservoir as those
of the fast (qS1) and slow (qS2) shear waves. Their most prone to causing casing damage, with episodic
Shear wave splitting in fractured reservoirs and resulting from earthquakes 455

Figure 15.38 Shear wave splitting analysed from AE events from siltstones in the Valhall field caprock. Event A exhibits the expected maxi-
mum attenuation of the slow wave S2. Event B exhibits maximum attenuation of the fast wave S1. Carter and Kendall, 2005.

microseismic stick-slip events close to the top of the reser- NE-SW strike (the dominant direction), or NW-SE
voir. Perhaps these ongoing bed-slip events were influ- strike. Note the consistent dip signs on both drawings,
encing the steep or vertical bed-limited jointing, giving despite the ‘vertical or sub-vertical’ assumption. These
stimulus to qS1 to qS2 reversals. The phenomenon dips, even if minimal, seem to be important for using
might possibly be related to the ‘90°-flips’ discussed earl- shear waves to distinguish between gas and oil in the frac-
ier in this chapter, which were suspected by the writer as tures, following Sayers, 2002b.
due to principal effective stress ‘flips’. Extensional and shear fractures were noted in each of
the principal strike directions. The dominant NE-SW
extensional set were continuous over hundreds of meters,
15.8.9 Shear-wave splitting and were responsible for the dominant permeability
and fluid identification at direction, as established by tracer tests. Along the crest of
the Natih field the shell-shaped 6  10 km antiform, which terminates
at a major fault zone, the bed-curvature being increased,
A series of interesting papers were published concerning there was evidence of the NW-SE striking extensional set
a large scale 3D shear-wave experiment performed over also participating in the fluid-flow network.
the 1 km deep fractured carbonate (chalky limestones) The scope of the nine-component three-dimensional
Natih field in Oman. The structure of the field was (9C3D) survey/experiment was impressive, with 10,800
described as ‘not very complex’ by Potters et al.,1999. 3C geophones, up to 1000 vibrator positions per day, and
The matrix permeability was a very low 1 to 30 mD. 22 million traces recorded on 2,000 tapes during the
Hard shales overlie the 300 m of carbonates, and these 32 days, and 28 km2 of field work. The survey basic
Fiqa shales exhibited some interesting anomalies as we grid size was only 25  25 m. Shear-wave anisotropy
shall see. The fractures in the carbonates were nearly exceeding 15% was registered over about half of the
vertical (a detail that seems to be important if both field. The average well flow rate in areas of large time
shear and normal compliances are to be involved, fol- delay was higher than that in areas with low S-wave
lowing Sayers 2002b, reviewed earlier). However the anisotropy, but significantly the local fracture swarms
authors of the two reports reviewed did not emphasise giving individual wells high productivity were too small
this aspect. to be detected seismically. Potters et al., 1999 suggested
Outcrop mapping included fractures too large to be that the absence of strong seismic anisotropy did not
generally detected by core or FMS analysis, meaning that however preclude the presence of fractured zones.
vertical wells were the usual poor samplers of typical Recalling the earlier critique of fracture density in this
sub-vertical structure. The outcrop, of necessity 50 km chapter, it is encouraging to note the authors’ reference to
distant, had joint character as shown in Figure 15.39a the ‘well-known ambiguity that a given amount of shear
(scale not given by Van der Kolk et al., 2001), but pos- anisotropy can be caused by an infinite number of combina-
sibly the same as the ‘1 m’ scale given for the reservoir in tions of fracture densities and sizes. Since different combina-
Figure 15.39b. The two fracture or joint sets had either tions have different rheological (and flow) properties, this
456 Rock quality, seismic velocity, attenuation and anisotropy

possible future geomechanics-based solution, since the


‘mechanical properties’ referred to are in the case of shear-
wave anisotropy, a function of the dynamic shear stiffness
(or compliance) for the case of vertical incident waves,
and a function of both the shear and normal dynamic
stiffnesses for the case of non-vertical waves (or non-verti-
cal fractures), following e.g. Sayers, 2002b.
As we shall indicate in Chapter 16, there are direct
links between the rock mechanics of pseudo-static rock
joint shear and normal stiffness behaviour, and their
stress-aperture behaviour, both hydraulic and mechani-
cal, with e  E. This is due to their common predic-
tion by the Barton-Bandis constitutive laws, using joint
characterization parameters JRC and JCS for wall rough-
(a) ness and wall strength respectively (Barton and Choubey,
1977, Bandis et al., 1981, 1983).
In Chapter 16 we will indicate the common mis-
match between the dynamic and static normal stiff-
nesses (roughly reflecting the differences between
dynamic and static moduli), which are therefore roughly
predictable. The dynamic shear stiffness (or compli-
ance) sensed by the slow shear-wave in the case of ver-
tical incidence and vertical fractures obviously also
carries information of relevance to the stress-closure-
aperture-permeability behaviour of the dominant
joints. By estimating the roughness parameter JRC
from core or well-bore images, the (effective, confined)
joint wall strength JCS could be estimated, which then
allows estimation of the mechanical aperture at the
given effective stress levels. Conversion to an estimate
(b) of hydraulic aperture is the final stage.
Figure 15.39 A partly contrasting, and partly consistent informa-
Van der Kolk et al., 2001, concluded the series of art-
tion from the surface exposure jointing (50 km dis- icles by Shell and their collaborators in Oman concerning
tant), and from the 1 km deep reservoir jointing at the Natih 3D shear-wave experiment, by presenting,
the Natih chalky-limestone reservoir in Oman. This reportedly for the first time, evidence that shear waves
was the site of a large shear-wave experiment in 1991. were sensitive to fluid type (gas or oil), in fractured or
Van der Kolk et al., 2001. jointed media. Regions of gas were characterized by slow
shear waves, and this had particular consequences for
two phenomena described by the authors.
problem needs to be solved possibly by incorporating a the- Firstly, the shear-wave splitting map of the Natih
oretical or empirical fracture size distribution’. The latter reservoir exhibited much larger splitting values (i.e.
sounds like a good solution, since the ubiquitous anisotropy) over the gas cap on the reservoir. This increase
microcracks seemingly favoured by Crampin, do not in anisotropy was due to the decrease in the slow shear-
seem to have utility in the face of the need (and exis- wave, which senses both the fractures and the fracture
tence) of the permeable fracture networks of interest filling fluid. A second phenomenon was the shear-wave
especially to oil companies. data from directly above the reservoir. The thick Fiqa
The authors also addressed another uncertainty. shale also exhibited a low shear-wave velocity anomaly,
‘Flow is strongly influenced by fracture aperture, a prop- but associated with a gas chimney.
erty which is, at best, only weakly expressed in the reser- Van der Kolk et al., 2001 used a semi-dynamic effec-
voir’s mechanical properties’. Here we may interject a tive medium model, to help to explain some of the
Shear wave splitting in fractured reservoirs and resulting from earthquakes 457

Figure 15.40 An effective medium BOSK model prediction of the shear-wave splitting magnitude for a set of 1 m long and 10 m long fractures,
as a function of fracture porosity, i.e. directly related to aperture. Van der Kolk et al., 2001. Note splitting %  (Vs1  Vs2)
Vs1  100.

observed phenomena at the Natih field. This model since they effectively define two different aspect ratios,
was based on a combination of the work of Budianski therefore influencing the assumed squirt-flow losses and
and O’Connell from the 1970s concerning the elastic the assumed stiffness.
moduli and visco-elastic properties of cracked fluid- The e  E inequality* is therefore a potential source of
saturated solids, together with extensions for arbitrary error for the case of effective medium modelling of rough-
crack orientation statistics from Sayers and Kachanov walled, tightly compressed (i.e. deeply buried) cracks,
in the 1990s. They termed the combined model the joints or fractures, when using only one aspect ratio.
BOSK theory. With this model, ‘squirt’ losses were also The crest of the Natih reservoir structure, with its
accounted for between the pores and fractures. This higher (50% or more) shear-wave splitting had higher
model correctly predicted that the slow shear-waves, local fracture densities than the flanks, but this was not a
polarized perpendicular to the fractures, were more sufficient reason for this higher value. Using the BOSK
influenced by the fluid type occupying the fractures, theory, van der Kolk et al., 2001 were able to investigate
than were the compressional-waves. (Classic Gassmann, the effect of fluid viscosity (causing dispersive, frequency-
1951 theory anticipates a sharp reduction in the P-wave dependent behaviour). Figure 15.41a shows the modelled
velocity in the presence of gas in porous media, and effect of gas replacing brine on the vertical P- and S-wave
supposes that the S-wave velocity is relatively unaffected velocities, as a function of fracture density.
by the type of fluid). The modelling results indicated that with a fracture
Figure 15.40 shows a BOSK model prediction of the porosity of 0.025%, the shear wave splitting was about
degree of shear-wave splitting as a function of fracture 50% higher in gas-filled, compared to brine-filled frac-
porosity, for a fixed fracture density of 0.2 (where e  tures. With lower fracture densities than 0.15, the ‘classic
N a3/V). To avoid the ambiguity of e, discussed at length
earlier in this chapter in connection with EDA, the
authors specified reservoir-like fracture lengths of a) 1 m * The two joint apertures will frequently differ by a factor of
and b) 10 m in these two BOSK realizations. Note that about 2 to 5, most for higher roughness JRC, and highest
there are subtle differences only where fracture porosity is normal stress level. This inequality was demonstrated at the
extremely small. The fracture porosity at Natih was about Technical University of Trondheim. Heimli, 1972, used a
0.1%. It is noted that the BOSK model, as for many other pre-instrumented intact core, subsequently split axially, so
effective medium models, (to be briefly reviewed soon), that different experimentally set values of E  0.05, 0.1,
0.2 mm could be known with certainty. These were
apparently makes no distinction between the mechanical
subsequently compared to the smaller hydraulic apertures (e)
(E  actual) assumed crack aperture and the hydraulic
back-calculated from flow tests. The e  E inequality has
aperture (e). In jointed rock (E) actually controls stiffness since been confirmed many times, and is explained by wall
and deformation moduli. In rock mechanics, we do not roughness effects, eventually quantified by JRC. See Figures
often use the crack aspect ratio. On the other hand, (e) 16.6 and 16.7 in the next chapter. Data sets for (e) and (E)
controls the intrinsic permeability, given as (e)2/12, and and an empirical JRC-based model are shown. (Barton, 1972,
both apertures (e and E) probably influence attenuation, Barton et al., 1985, Barton and Quadros, 1997).
458 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.42 The ratio of seismic Qgas/Qbrine in modelled fractured


media, using the BOSK model. Shear waves were
much more attenuated than compressional waves by
the introduction of gas. Van der Kolk et al., 2001.

The authors concluded that vertically moving P-waves,


and the fast shear wave, were hardly affected by whether it
was gas or brine in the fractures. It was the slow S-wave
that was polarized perpendicular to the fractures that
registered the fluid type, i.e. whether of low or high
compressibility.
The BOSK model was also used by van der Kolk et al.,
2001 for studying the effect of frequency, and the relative
effects on attenuation of gas or brine. The ratio of seis-
mic Q for gas compared to the seismic Q for brine
(Q gas/Q brine) indicated greater attenuation of the
S-waves with the introduction of brine. This result is
shown in Figure 15.42.
Figure 15.41 (a) A BOSK effective medium prediction of the effect
The authors made some important and undoubtedly
on vertical P- and S-wave velocities, of gas replacing correct conclusions, which are nevertheless contrary to
brine. (Note Vs refers here to Vs2, the slow shear wave standard exploration practices, and will therefore be
polarized perpendicular to the fractures.) Two sets of quoted in full: ‘Both the observations and the (BOSK)
vertical fractures, in an isotropic, low porosity matrix modelling suggest that the S2 shear-wave propagation
were modelled, with a wide range of fracture densities. depends on the fluid type in the fractures. For propagation
(b) P-wave and polarized S-wave surfaces for gas and parallel to and polarization perpendicular to the plane of
brine filled fractures, showing the modelled influence the fractures, the observed effect is exactly opposite to what
of wave direction relative to two sets of vertical frac- is predicted by Gassmann fluid substitution in a porous
tures. The fractures are oriented 45° to either side of matrix. We can conclude therefore that Gassmann theory is
these figure axes. Van der Kolk et al., 2001.
not sufficient to model fluid replacement in (heavily) frac-
tured media. (Matrix porosity effects must still be included
in the calculations, of course). Many standard exploration
behaviour’ of P-waves more affected by fluid type than practices are based on Gassmann substitution, e.g. direct
S-waves was seen. When P-waves crossed the fractures, hydrocarbon indicators (DHIs) such as bright spots, flat spots
rather than paralleling them, the predicted sensitivity and sometimes AVO effects. The results obtained above
to fluid type was found to be stronger in P-waves than suggest that these techniques may be invalid in fractured
S-waves, agreeing with the traditional behaviour. media and new extensions should be explored ’.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 459

or 1/Qseis was derived from comparison of spectra


recorded for different azimuths, but with identical
source-receiver offsets. Absolute attenuation is typically
recorded by comparison of spectra recorded along an
individual ray path. The authors utilised the field meas-
urements of travel time anisotropy, and were able to
invert for fracture strike, fracture intensity (usually
ambiguous), and scale length (presumably making the
intensity formulation non-ambiguous).
When attempting to measure the absolute values of
attenuation with the traditional spectral ratio and the
instant frequency method of Dasios et al., 2001, they
found that the instantaneous frequencies of the events
fluctuated strongly: the heterogeneous rock mass giving
zones of apparent negative Qseis. The authors argued that
since the fast and slow shear-waves had similar wave
lengths, they therefore probably sampled the same het-
(a) erogeneities. They therefore analysed the differences in
the instantaneous frequencies of the two waves. Despite
the scatter, the differences in the two sets of frequencies
were consistent with their location in either the over-
burden or in the fractured reservoir.
Figure 15.43b, showing these differences in instanta-
neous frequency, indicates increased attenuation of the
slow shear wave, which had lower frequency in the reser-
voir layer (5,800 to 6,200 ft), but essentially the same
frequencies in the overburden. One could perhaps spec-
ulate on another less dominant fracture direction in the
overburden, in view of the mostly not quite zero differ-
ences in instantaneous frequencies.

(b) 15.9 Dual-porosity poro-elastic


modelling of dispersion and
Figure 15.43 (a) Rose diagram showing the consistent P-wave
fracture size effects
anisotropy directions in the reservoir (black) and the
overburden (grey). (b) Differences in frequency
between the fast and slow shear-waves, indicating a
The ability to model various aspects of jointed rock behav-
consistently lower frequency for the slow shear iour has existed for many years, and is a complex and con-
waves in the fractured reservoir from 5,800 to 6,200 stantly expanding field. No attempt can therefore be made
feet. Maultzsch et al., 2005. to give an exhaustive treatment in a single section of a sin-
gle chapter. Since this section will address dual porosity
poro-elastic modelling, with several examples, we will first
At the end of Chapter 14, an analysis of the Claire Field summarize the different modelling needs and capabilities
by Maultzsch et al., 2005 concerned P-wave attenua- developed in rock mechanics, which is an increasingly close
tion anisotropy derived from analysis of field VSP data. neighbour of geophysics, particularly in recent years when
In this case, with no gas cloud, the fractured reservoir micro-deformation depending on joint stiffnesses, and
displayed much stronger anisotropy than the overbur- micro-flow simulation depending on apertures, has
den (Figure 15.43a), the authors drew attention to the become important for interpreting seismic response.
advantages of using relative attenuation, rather than It is probably fair to say that rock mechanics modelling
absolute attenuation. Their concept of relative attenuation advanced very far in the last several decades of the 20th
460 Rock quality, seismic velocity, attenuation and anisotropy

century, for the dual reasons of rock engineering and civil subsequently replaced by simply deformable, then fully
engineering needs, and due to the relative accessibility of deformable blocks in UDEC, with the ability to model
input data, specifically from drill core, shallow exploration dynamic loading, micro- or macro-deformations, and
adits, and extensive use of shallow refraction seismic, plus fluid flow within the joints, but not in the matrix,
sampling from within tunnels during construction. This where only pore pressures were modelled. Here we see
has been a ‘constant’ for many decades, although descrip- one of the limitations, which of course is no longer there
tion (and use) of the joint properties has improved dramat- when only modelling flow in (un-jointed) porous media.
ically, with perhaps less isotropic continuum modelling UDEC first had Mohr Coulomb linear joint strengths
than previously. and linear stiffnesses, then a Cundall continuously yield-
In contrast, geophysics data acquisition abilities ing law, followed by non-linear shear strength and stress
(and quantities of data) have exploded in just the last dependent shear and normal stiffnesses, following the
decade or two, with digital recording of increasingly com- block-size-sensitive Barton-Bandis constitutive model, in
plex 3D 4C receiver arrays, both on-land and especially which most of the input data can be derived from simple
off-shore. This development continues unabated. There is index tests performed on drill-core, giving joint parame-
now an enormously increased need for realistic numerical ters JRC, JCS, and r. Block-size determined the scaling
models for geophysical interpretation, particularly in the of these two joint roughness and compressive strength
area of structural anisotropy, fractures sizes and properties, parameters, and the core-logged Q-value gave a stress-
and the frequency dependence of their seismic responses. dependent deformation modulus. As we have seen in Part
I, (and in Figure 15.33) the P-wave velocity from shallow
refraction seismic could be used to estimate the static
15.9.1 A brief survey of rock deformation modulus, and to extrapolate or interpolate
mechanics pseudo-static such data from borehole to borehole.
models of jointed rock This non-linear 2D model was termed UDEC-BB.
Flow and fluid pressures were modelled by converting
By way of an extremely brief history, one can mention the the physical joint apertures (E) developed at any time
finite element modelling with joint elements that was during the modelled joint deformation, into the (usu-
developed at the end of the 1960s and promoted for rock ally) smaller hydraulic apertures (e), using a linear-
engineering use by Goodman et al., 1968 and others. laminar flow assumption with joint permeability given
These authors, and Goodman, 1970 were perhaps the by k  e2/12. The conversion between (E) and (e) was
first to define the (pseudo-static) joint normal and shear found to depend on roughness JRC (Figure 16.7).
stiffnesses needed for input to their 1D joint-slip elements Subsequently 3DEC was developed by Cundall,
for discontinuous finite element modelling. The deforma- together with colleagues at Itasca, giving the ability to
tion and stability of slopes, dam foundations and tunnels model three-dimensional assemblages of jointed,
were the primary focus in rock engineering, and fluid deformable blocks, with any desired moduli and joint or
flow and effective stress analyses were of course required fault properties. Linear strength and stiffness laws were
too. These early 2D FEM models had deformable joint followed for the joint sets and faults, to reduce calculation
elements requiring linear estimates of normal and shear time. More recently, the pore space defined by the com-
stiffness and of course frictional and eventual cohesive plex, three-dimensionally deforming, intersecting joint
strength. Three dimensional FEM modelling, for exam- sets was fully defined, allowing flow modelling and
ple for dam-and-foundation interaction studies, with dynamic effective stress modelling to be performed, also
more limited numbers of (major) joint and fault planes in three dimensions. Of course there are more dedicated
was also performed during the 1970s and subsequently, 3D fracture flow models like FRACMAN (Dershowitz/
but was obviously very time-consuming. Golders) and NAPSAC (AEA Harwell), but these are
In the early 1970s Cundall, 1971 developed a 2D finite lacking comprehensive joint deformation modelling.
difference distinct element model for randomly or regu- There are now several numerical models in use in
larly jointed rock masses, making it easier to model rock mechanics for also modelling rock failure or crack-
large assemblages of blocks, with the ability, if required, to ing of the matrix blocks, caused by over-stress. These
follow large deformations by tracking edge and corner can accommodate a more limited number of pre-existing
contacts. Initial rigid block calculations in
DEC were joints. Prominent among these are the particle flow codes
Shear wave splitting in fractured reservoirs and resulting from earthquakes 461

of Cundall (PFC2D and PFC3D), and the non-linear 15.9.2.1 Schoenberg slip interface concept
fracture mechanics code of Shen (FRACOD). The lat- Schoenberg 1980 modelled elastic wave behaviour using
ter in particular, seems capable of realistic modelling of linear slip interfaces. These allow reflection, transmission,
log-spiral type break-out around tunnels or boreholes, conversion, and delay to take place at the modelled inter-
without recourse to ‘manual’ degradation of cohesion, face, with the magnitudes depending on the specific stiff-
and mobilization of friction, as needed when using an ness, the frequency content, and the angle of incidence.
inappropriate ‘c  tan ’ formulation. This linear Mohr The assumption is that when an elastic wave propagates
Coulomb law is acceptable, or at least much used, for across a fracture, there is a displacement discontinuity
particulate clays and sand modelling in soil mechanics, that is linearly related to the normal or shear force gener-
and is reasonable for pre-existing planar joints, though ated. The seismic particle displacement is discontinuous,
probably best for faults. It is incorrect for previously while the seismic stresses are assumed to be continuous.
intact hard rock, where ‘c then tan ’ is a more correct In the linear slip model, the displacement discontinuity
formulation than ‘c  tan ’, due to the widely differ- vector u was assumed to be linearly related to the trac-
ent strains involved in cohesive failure of a brittle tion t as follows:
material and subsequent frictional sliding along the fail-
ure surfaces. (See discussion by Barton, 2004b).
 ZT 0 0 
 
u   0 ZN 0 t (15.7)
 0 0 ZT 
15.9.2 A very brief review of slip-  
interface, fracture network
and poro-elastic crack
In geophysics it became customary to talk of fracture
models
compliances, with inverted nomenclature and units (e.g.
MPa/mm for stiffnesses, and m.Pa1 for compliances).
Concerning the dynamic modelling of cracked rock in
Naturally, the compliances or stiffnesses used in geo-
geophysics, we can quote Tod, 2002 who is a prominent
physics refer to the dynamic properties of the joints or
new contributor to this field: ‘There are many theories
fractures, which generally have somewhat greater stiff-
available in the literature, resulting from a range of theoret-
ness (or lower compliance) than the pseudo-static values
ical backgrounds that provide a description of an effective
commonly used in rock mechanics modelling. These
medium appropriate to describing the properties of a matrix
differences have been mentioned many times, and will
material permeated with cracks on a length scale far less than
be quantified further in Chapter 16.
the wavelength of seismic waves. While many of these theories
agree qualitatively and are capable of describing a number of
observed features, each has its particular shortcomings.’
For the modelling of cracks in geophysics, we will not 15.9.2.2 Hudson effective medium concept
go further back in history than to mention Schoenberg Hudson, 1980, in contrast to Schoenberg, utilised a
1980, and Hudson, 1980. These classic developments ‘method of smoothing’ for the effect of modelled cracks,
assumed, for greater simplicity (there was enough math- which was capable of representing the elastic parameters
ematics without fluids), that there was no exchange of of a ‘cracked’ material, in the form of an effective
fluid, either between the fractures or cracks themselves, medium. This allowed calculation of the effect of inci-
or between these and the rock matrix. The importance dent dynamic waves of long wavelength. Subsequently,
of fluid had of course been known for a long time, and Hudson et al., 1996 extended the model to allow for
was modelled within the pore space by Gassman 1951 the cracks to be connected via the porosity of a rock
and within the cracks by O’Connel and Budiansky, matrix. In this extended case, cracks could be deformed
1977, and by many others since then. Thomsen, 1995, by an incident wave in a manner that depended on
showed how seismic anisotropy was enhanced by trans- their aspect ratio and on their orientation with respect
fer of fluid between fractures and equant porosity, with to the incident wave. Clearly the modelling of intrinsic
perfect pressure equalization at very low frequency, and attenuation mechanisms such as squirt flow, and its fre-
reduced equalization at higher frequencies, giving the quency dependence, was transformed by this extended
‘unrelaxed’ behaviour modes. capability.
462 Rock quality, seismic velocity, attenuation and anisotropy

expected, as isolated (stiff ) features, while at low fre-


quencies they behaved as ‘poorly-drained’, since squirt
losses at lower frequencies cause a ‘soft’ behaviour
because of some (micro) drainage.
The model developed by Tod effectively allows the
crack density to decay with an increase in applied stress,
from an initial value representing the unstressed state.
However, the model is purely elastic, and relaxes to
its original state during unloading. Crack density was
designed to decrease with increasing compressive stress,
from e.g. 0.1 at zero effective stress to e.g. 0.05 at
200 MPa. Only cracks with normals lying in the h min-
imum direction were assumed to remain open.
The model is capable of capturing the changed
anisotropy caused by fluid pressures and applied stresses,
which impact the aspect ratios of the cracks. However the
pore space, illustrated in Figure 15.44a is non-compliant.
Wave speeds approach that of the matrix at high pres-
sures. Shear waves proved to be more sensitive to pressure
change than the compressional waves, as known from
other studies.
The non-compliant pores of the Tod model, transfer
fluid to the physically unconnected cracks, therefore giv-
ing dispersive, or frequency-dependent velocities and
attenuation. The theory predicts that there is a pressure
at which anisotropy reaches a maximum value, before
the conducting properties are reduced, by increasing
stress, eventually to that of the uncracked matrix. Tod
also extended his modelling to the case of faults with
parallel cracks aligned at an angle to the main fault plane
(Figure 15.44c).
Since these authors, and also Chapman, 2002 and
2003 (see later), utilise aspect ratios that can in practice
be reduced to very low values by high stress, the
inequality of physical (E) and hydraulic (e) apertures
will affect interpretation, since fluid volumes in actual
cracks or joints are greater than hydraulic capacity, since
permeability k  e2/12. In dynamic modelling seen so
Figure 15.44 (a) Schematic of the pore space created by the misfit
of (sand) particles, (b) a possible distribution of cracks
far, this distinction does not appear to have been treated.
that are subject to a sub-parallel stress, c) oblique However, the improved model described by Tod et al.,
cracks in a fault zone applied in a new model. Tod, 2002, (together with Cambridge Professor ‘seismic’
2002, and Tod et al., 2002. Hudson: there is a prominent Professor J.A. Hudson in
rock mechanics too), provides a new level of sophistica-
tion, since crack aspect ratio, crack density, crack orien-
15.9.2.3 Tod crack density decay model tation and responses to applied stress and fluid pressures,
Tod et al., 2002, extended this model further, by allowing are each incorporated into the permeable, but non-
for a continuous but independent distribution of both compliant pore space between the matrix ‘grains’. The
crack orientation and aspect ratios, and by allowing new formulations and the interdependencies of crack
each to depend on the applied stress and on the fluid orientation, aspect ratio and applied (effective) stress
pressure. At high frequencies, the cracks behaved, as gives an elastic loading and unloading behaviour, but
Shear wave splitting in fractured reservoirs and resulting from earthquakes 463

(a) (b) (c)

(a) (b) (c)

Figure 15.45 Three examples of the nucleating branching fracture networks, showing their changing attenuation and frequency depend-
ence. Source wavelets of 30 Hz applied at centre of each model. Vlastos et al., 2003a. Reproduced by kind permission.

with no hysteresis, and no stress history, in obvious but does not include squirt or intrinsic attenuation
contrast to the larger-deformation modelling of the losses. Nevertheless, scattering attenuation is also found
rock mechanics UDEC-BB code. to be frequency dependent, showing Qseismic values as
low as 1/0.6  1.7 for case (b) in Figure 15.45.
In the numerical fracture models illustrated in Figure
15.9.2.4 Vlastos-Narteau automaton model 15.45, background values of Vp and Vs were 3.3 km/s
Interesting new developments in modelling capabilities and 2.0 km/s respectively. Density was 2.2 kg/m3.
in geophysics were also demonstrated by Vlastos et al., Vlastos et al., 2002 and Vlastos et al., 2003a used what are
2002 and Vlastos et al., 2003a. The first authors considered by the writer to be unrealistically equal normal
Vlastos and Narteau, utilised the 2D finite difference and shear compliances (ZN  ZT  5.6  1010 GPa1:
method termed the multi-scale ‘cellular automaton presumably GPa1.m?), which (may) correspond to rock
model’ developed by Narteau, 2001 to study progres- mechanics normal and shear stiffnesses (Kn and Ks) of
sive changes of attenuation in nucleating, growing, about 1.8 MPa/mm. In general terms, due to the higher
branching and coalescing arrays of fractures. They stud- dynamic modulus of rock masses, it is likely that the
ied the effects of scattering attenuation in what evolves, micro-deformation, dynamic compliances should be
through shear fracturing, into an anisotropically frac- much lower (i.e. much stiffer in rock mechanics termi-
tured network, with fractures of increasing length and nology) than the macro-deformation values typically
reduced frequency as shear localization develops. measured in ‘static’ loading tests on joints, as frequently
Examples of the successive stages of fracturing that performed in rock mechanics, and the shear compliance
can be generated are shown in Figure 15.45. These are should perhaps be higher (i.e. the shear stiffness lower).
successive ‘snapshots’ generated by a particular realization Presumably with rock mechanics experiences of e.g.
of this dynamic network model. Ks/Kn  1/10 or an approximated Z␶/ZN  10, even
In the seismic attenuation modelling described by the greater attenuation, and greater anisotropy would have
authors, only the scattering attenuation component is been indicated. Clearly there is promise for future links
modelled. This is related to the structural hetero- between fracture frequency, fracture character and seis-
geneities (i.e., the growth and coalescence of fractures), mic attenuation measurements.
464 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.46 ‘Snapshots’ of the anisotropic, elliptical, pore pressure propagation with time, in the Vlastos et al., 2003b, fractured network model.
Reproduced by kind permission.

Vlastos et al., 2005 presented a more comprehensive into a simulated 2560  2560 m model. They show
version of the Narteau ‘stochastic-deterministic’ frac- interesting ‘snapshots’ of the anisotropic, elliptical pore
ture network modelling, with a 12-stage evolution of pressure progression that develops with time.
fracturing. Application of the spectral ratio method to Figure 15.46 reproduced in grey-scale, shows the pore
the scattering attenuation, resulted in increasing then pressure development 10, 40, 70, and 100 hours after the
declining attenuation, giving seismic Q values reducing start of injection. They explained that the ellipticity of the
from 100 to 7 to 2.5 (with the maximum density of pore pressure field, besides being due to the presence of
fracturing), and increasing again through 6, 10 and 100 fracturing, was due to assuming a diffusivity along the
as the fractures were fewer and longer, as typified by fractures, many orders of magnitude greater than the
stage c) of Figure 15.45. background.
It is clear that the wide variety of fracture-scale lengths As the authors pointed out, injection decreased the
illustrated in the most well-connected networks were effective stress, thereby increasing the compliance (or
within the so-called ‘percolation threshold’, and were reducing the stiffness) of the modelled fractures. On this
largely responsible for the low Q seis values in these ‘cen- occasion a Schoenberg, 2002, non-linear compliance
tral’ models, typified by case b) in Figure 15.45. formulation was used. An increased fracture opening
Vlastos et al., 2003b, with co-authors from BGS and and/or pore throat size, could have effectively decreased
the universities of Berkeley, Edinburgh and Cergy- the stiffness of the rock and rock mass in terms of squirt
Pontoise in France, and Lin et al., 2004, used a 2D fluid flow, thereby reducing velocities and increasing intrin-
flow model and a pre-existing, anisotropically fractured sic attenuation.
numerical network, similar to the above, to study the The authors came to the surprising conclusion that
seismic signatures of a central (borehole) fluid injection the P-waves were not sensitive to pore pressure changes,
Shear wave splitting in fractured reservoirs and resulting from earthquakes 465

as opposed to the S-waves and coda waves that showed 15.9.2.5 Chapman triple-porosity poro-elastic
high sensitivity. Could the reasons for this be found in model
their assumptions for compliances, in relation to rock Chapman, 2002 and 2003, developed a dual-porosity
mechanics experiences of the approximate inverse: the poro-elastic model, based on the following important
stiffnesses? The authors reportedly set compliances as observation that typical laboratory samples, clearly
follows (but presumably the unit should be Pa1.m, as unfractured, nevertheless display dispersion, anisotropy,
otherwise the stiffnesses would be many orders of mag- stress sensitivity, and dependence on fluid type and
nitude too high): degree of saturation, as we have seen in numerous con-
texts in earlier chapters. Chapman’s argument for develop-
At zero stress: Z(To)  5.681  109 GPa1 ing his new model was that, when the fractures or
cracks were removed from preceding models, a linear-
Z(No)  2.8409  1010 GPa1 elastic material remained, contrary to observation.
Chapman et al., 2001 contrasted his more recent
If we assume a typing error, and intended units of poro-elastic double (actually triple) -porosity model
Pa1.m, then rock mechanics stiffnesses, if they had with the anisotropic poro-elastic model of Zatsepin and
been equal to the inverse, would have been: Crampin, 1997. Although the driving process of this
earlier model was the migration of fluid along inter-
Ks  0.176 MPa/mm crack pressure gradients (and subsequent preferential
crack closing), the calculation of the induced velocity
Kn  3.52 MPa/mm changes actually relied on crack models which assumed
that fluid could not move at the time scale of a seismic
These give a recognizable ratio Kn/Ks of 20. However, wave. As pointed out, this could be a good approxima-
the equivalent Kn value, as will be seen Chapter 16, is tion for high frequency laboratory ultrasonic experi-
exceptionally low, and the shear stiffness also low. Both ments, but not for low frequency e.g. 100 Hz field data.
are representative of an effective normal stress of per- The restriction that fluid should not move also ruled
haps less than 0.1 MPa. This could be the explanation out the (correct) modelling of attenuation.
of the lack of P-wave sensitivity to pore pressure Chapman had earlier presented a poroelastic theory
change, since the modelled rock mass was, perhaps that modelled the effects of squirt flow at the grain
inadvertently, given the equivalent of a high confining scale, which gave a good match to experimental data.
stress. At ‘infinite’ stress, the authors used Z(To)/5 (i.e. Chapman’s new theory combined two or more length
Ks  5) and Z(No)/2 (i.e. Kn  2), changes that would scales explicitly: the grain scale and a set of fractures of
be greatly exceeded in rock mechanics terms, when the any desired characteristic length (e.g. stress-aligned
original stiffnesses were so low, due to the assumed low dominant jointing of many potential scales: 0.01, 0.1, 1,
stress. 10 metres). He termed the latter ‘meso-scale’ anisotropic
Real rock joints display strongly non-linear pseudo- fractures. Naturally, he also modelled the all-important
static stiffness at lower stress levels, when subjected to interaction between this meso-scale and the grain-scale
mechanical loading causing macro-deformation. Further- equant matrix porosity and ellipsoidal microcracks.
more, near-surface rock masses clearly display a strong Because there may be three or more scales involved, it
dependence of P-wave velocity on effective stress level has been termed a ‘triple-porosity’ model in this book.
(and therefore pore pressure), at least in the first several An important feature introduced by the inclusion of
hundreds of metres, and sometimes to depths of one or fractures was that dispersion occurred at lower frequency
even two kilometres if the rock has high modulus, or ranges than those over which the micro-structure
if there is over-pressure. The rock quality Qc -Vp - dispersion occurred. He found that the larger the size of
porosity – depth relations which are empirically based the fractures, the lower was the relevant frequency band.
(e.g. Figure 15.33 and Chapter 15), confirm the relevance Figure 15.47 gives a clear demonstration of the abil-
of low Kn at low stress, with a strong stiffening at high ity of this model of equant porosity with elliptical
stress. So the authors’ conclusion regarding Vp not sen- microcracks (first without meso-scale fractures), to
sitive to pore pressure changes, may inadvertently be a model frequency dependent velocities (dispersion) in the
function of the ‘equivalent-to-low-stress’ input data, and case of an isotropic model without fractures. Figure
the application of high (fracture-closing) stress levels. 15.47 a and b, show that the dispersion is predicted
466 Rock quality, seismic velocity, attenuation and anisotropy

Next, Chapman added fractures of 10 cm length, at a


density of 0.05, making the model transversely isotropic.
Figure 15.47c shows the dispersion of the qP-wave veloc-
ities travelling parallel and perpendicular to the model
fractures. Clearly, the parallel waves are unaffected by the
fractures, but squirt flow gives dispersion for the perpen-
dicular waves. There is a significant reduction in velocity
at low frequencies, but no reduction in velocity above
about 1 kHz, due to the apparent stiffness of the fluid
inclusions. Introduction of even small 10 cm long frac-
tures had caused dispersion to begin at lower frequencies.
Figure 15.48 shows modelled shear wave velocity
anisotropy [defined as before, as 100  (qS1 – qS2)/ qS1]
for the case of a mix of microcracks of 200
m size, 10%
equant porosity, and meso-fractures of up to 1 m size.
(See visualization of the model in Figure 15.48a).
Shear wave propagation was at 20° incidence to the
fractures. Anisotropy reduced with increasing frequency.
In contrast, Figure 15.48b had no equant porosity (0%)
and crack density is reduced from 0.1 to 0.02. A marked
reduction (and smaller variation) of shear wave
anisotropy can be noted. The absence of the pores, means
that fluid can no longer flow out of the microcracks with
the passage of the seismic waves, so anisotropy and atten-
uation are reduced. Note that the shear-wave anisotropy
is only weakly following the ‘100e’ rule-of-thumb of
Crampin, since the two examples should then show
10% and 2% respectively.
When modelling the attenuation, there was a singu-
larity (zero attenuation) in the case of the pure shear wave,
when the frequency was only 40 Hz, and the angle of
incidence was 90°. This was believed to be due to the
perfectly aligned fractures† this aspect being avoided in
the new Tod, 2003 model, which utilises a ‘nearly
aligned’ crack set.
When Chapman modelled with a much higher fre-
quency (3 kHz), the qP, qS and pure shear S waves were


On the subject of modelling P-wave anisotropy in the
case of perfectly aligned fractures, Willis and Rao, 2005
Figure 15.47 (a) (b) Chapman model of dispersion with no meso-
pointed out that a P-wave can be cancelled out due to the
fractures. (c) dispersion and anisotropy when 10 cm
time delay of the wave, when travelling between two
fractures are added. Chapman, 2003.
fractures, thus creating a significant notch (or even a null) in
the spectral ratios of reflected seismic traces. This happens
when the P-wave length is about twice the fracture spacing.
over a limited frequency band, roughly between 1 kHz However, they also suggest that the frequency location of the
and 100 kHz. The particular model had moduli of notches themselves could be used to determine the fracture
17.5 GPa, density 2300 kg/m3, a microcrack density of spacing and losses, due to scattering attenuation from the
0.1 and 10% porosity. fractures.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 467

(b)

(a) (c)

Figure 15.48 (a) Visualization of Chapman’s model of connected equant pores, microcracks and fractures with low aspect ratios, (b) S-wave
anisotropy with 10% porosity, and ellipsoidal microcracks and fractures of various sizes, giving a crack density of 0.1. (c) simi-
lar to (b) but with 0% porosity, and 0.02 crack density, which caused reduced attenuation, and reduced S-wave anisotropy.
Chapman, 2003.

all attenuated, with (1000/Q) in the range 18 to 24 roughness (JRC), the usual inequality: E  e occurs,
(approx.), i.e. QqP and QqS values in the range 42 to 56, with e  E only when roughness is absent, and JRC  1.
which is typical for Qp and Qs of moderately fractured In reality, a chosen aspect ratio will have less permeabil-
rock closer to the surface (see Chapter 10). ity than assumed, due to joint roughness.
This brings one to the interesting question of stress Because of varying degrees of roughness for most
sensitivity. By setting smaller aspect ratios for the meso- joint types, and the usual inequality of these two
fractures compared to the microcracks, one can in some apertures, dispersion will likely begin at lower frequen-
way mimic the effect of higher stress, and thereby cause cies than might be modelled; there will be extra resist-
less attenuation (higher Qseis), as the fluid cannot ‘squirt’ ance for fluid flow (in relation to the modelled aspect
or flow so easily, in response to the passage of the seis- ratio), as roughness increases. Yet it will contain the
mic waves. assumed volume of fluid. The significant inequality of
Setting a certain aspect ratio is like creating a phys- E and e is demonstrated in the explanation of some
ical aperture (E), whereas the assumed permeability of rock mass ‘groutability’ terminology, in Figure 15.49,
this crack of ‘fixed’ aspect ratio is actually a function of and by UDEC-BB modelling of tunnels, in Figure
its hydraulic aperture (e). Because of joint or fracture 15.50.
468 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.49 Physical demonstration of joint aperture concepts (E) and (e) regarding groutability. Barton and Quadros, 2003.

15.9.2.6 Maultzsch-Chapman fracture size where a  crack radius, and N/V is a number per unit
estimation volume (  ‘number density’).
Maultzsch et al., 2003, with colleague Chapman from the This means, as they point out in Figure 15.51a, that
Edinburgh BGS Anisotropy Project, broached the ques- the same crack density can be caused by a few large frac-
tion that seems to represent an important ambiguity in tures or many small cracks. For instance, a fracture dens-
many earlier models in geophysics, namely how best to ity of 0.05 means that N/V (where e  Na3/V) can be
interpret the commonly used ‘crack density’ formula: expressed in many different ways, such as:

Na 3 a) 50  1 mm cracks/1 cm cube
e  ga 3 or e  (15.8)
V b) 5  1 cm cracks/10 cm cube
Shear wave splitting in fractured reservoirs and resulting from earthquakes 469

Figure 15.50 The inequality of physical aperture (E) and hydraulic aperture (e), demonstrated in a distinct element UDEC-BB model of
tunnels in Oslo. Makurat and Barton, 1988, NGI contract report. Makurat et al., 1990a.

during field logging, namely attempt to present the spac-


ing statistics of the different sets of joints (with
a mean m1 for each set), and where possible, also pres-
ent the length statistics of the joints belonging to the dif-
ferent sets, and whether both or only one end of the joint
in question is visible in the exposure. (ISRM, 1978).
Quantification as joint spacing and length and num-
ber of sets, as used by engineering geologists, would be
Figure 15.51 The problem of the crack density definition e  the theoretical ideal, but with compromise necessary
Na3/V used in geophysics is that both the illustrated
due to the ‘invisible’ nature of the targets in geophysics.
representations of fractures of different size can have
Mixed scale fracturing obviously needs to be modelled
the same numerical value of e. From Maultzsch et al.,
2003.
without ambiguity concerning fracture size, since it
is of fundamental importance when attempting to
assess fractured reservoir structure, as also argued by
c) 50 10 cm fractures/1 m cube
van der Kolk et al., 2001 in the case of the Natih field
d) 1  1 m fracture/2.7 m cube
in Oman.
When an engineering geologist is logging exposed
While the crack density can be roughly estimated rock masses or core, he also uses one or more rock mass
from the time delay of split shear waves, since a shear classification methods, such as RQD, Q, and RMR).‡
wave anisotropy of approximately 100e is expected Naturally, being able to see the rock surface exposure,
according to Crampin, the result is seemingly ambigu- tunnel wall or core, is obviously of inestimable value for
ous, since fracture compliances will also play a role in giving this more correct description of reality. The
the time delay and anisotropy estimate. This ambiguity microcrack scale is of course excluded from general
is hardly satisfactory when trying to understand the field observations.
structure of a fractured reservoir. Maultzsch et al., 2003, investigated the length-scale
This geophysics crack density concept is in stark con- frequency-dependence of the Chapman model, taking
trast to what engineering geologists are expected to do account of fractures up to 10 m size in Figure 15.52.


The first two terms of the rock mass quality Q-estimate are one for rock mass stability and inter-connectivity to fluids. It is
step simpler than estimating m1 for each set, as above. For tempered by the next pair of parameters Jr/Ja, which closely
RQD, one records the percentage of core pieces longer than represent the coefficient of friction, with contrasting
10 cm, dividing this by the rating (Jn), for the number of joint contributions from joint roughness and clay filling, in the first
sets (see Appendix A). These first terms in the Q-value estimate instance for the least favourable joint set. The combined value
(RQD/Jn), give a close approximation to relative block size, and Q  RQD/Jn  Jr/Ja, often termed Q-prime, clearly has
the degree of freedom for block movement. This is fundamental strong links to seismic attenuation and anisotropy.
470 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.52 A further example of the scale and frequency depend-


ence of shear wave splitting anisotropy, according to
the Chapman model, that connects pores, microcracks
and formation-scale fractures. In this example, from
Maultzsch et al., 2003, the waves are propagating at
30° to the fracture planes, and are split to give the dis-
persive anisotropies shown.

Here the percentage anisotropy of the polarized qS1


and qS2 shear wave velocities is shown. It is implied
that shear wave splitting anisotropy can be registered
from wave splitting at numerous scales of crack/fracture
sizes. However it is understood that attenuation in a Figure 15.53 Comparison of Chapman model to the physical
well fractured reservoir may prejudice the arrival of the saturated sandstone-and-fracture model of Rathore
slow qS2 wave. et al., 1995. From Maultzsch et al., 2003.
For any given fracture radius, the shear-wave
anisotropy is seen to decrease as frequency increases. The
larger the size of fractures, the lower the frequency disturbed rock masses, such as along the instrumented
range where velocity dispersion and frequency depend- sections of the San Andreas Fault, as reviewed earlier in
ence of anisotropy occurs. This property was subse- this chapter.
quently used by the authors for inversion to fracture size. Maultzsch et al. 2003 demonstrated the application
As also in Figure 15.48, the modelled results of shear of the Chapman model with two examples. One was a
wave splitting shown in Figure 15.52 suggest that the novel physical model of known fractures and pore
EDA concept with focus only on microcracks as the structure located in IKU, Trondheim, Norway, made
source of shear-wave splitting, may need to be re- by embedding thin metal discs into a sand-epoxy
evaluated. Microcracks, as modelled, appear to give a matrix, with subsequent acid leaching of the discs to
‘constant’ potential source of shear-wave splitting create cracks of known geometry and orientation
independent of frequency. In the examples shown in (Rathore et al., 1995). This physical model will be
the two figures, anisotropies of 7.5%, 3.5% and 4.25% described in greater detail at the end of this chapter. It
are shown. The ‘meso-fractures’ of 1 m and 10 m appears likely to have aperture characteristics satisfying
(radius) apparently have an equally strong role in shear- E  e, (i.e. without contact or apparent stress transfer,
wave anisotropy as microcracks, but only at the seismic and with limited, if any, roughness). The Maultzsch et
frequencies appropriate to earthquake studies. There is al., 2003 match to these laboratory determined qP, qS
therefore no good reason for only assuming that microc- and Sh velocity components with azimuth, was excel-
racks are responsible for shear-wave splitting in seismically lent, and is shown in Figure 15.53.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 471

15.9.3 Applications of Chapman


model to Bluebell Altamont
fractured gas reservoir

Since the Chapman, 2002 and 2003 dynamic equivalent


medium model handles squirt flow in a triple-porosity
poroelastic medium with porosity, microcracks and a
fracture set of any desired fracture length or density, it
is logical that it becomes a useful tool for demonstrat-
ing both the frequency dependence and azimuthal vari-
ation of the attenuation (1/Qp) of P-waves.
Maultzsch et al., 2002 used the Chapman model,
and various fracture sizes to demonstrate azimuthal
dependent Qseis (strictly 1/Qp) and also frequency
dependence, each for four different fracture sizes. This
synthetic model had a common fracture density of 0.05,
and fracture sizes (radii) of 1 mm, 1 cm, 10 cm, and
1 m. Figure 15.54 shows the predicted results. Note the
tendency for 10 cm and 1 m fractures to give highest
attenuation at lower seismic frequencies.
Maultzsch et al., 2003 applied the Chapman model
to nine-component VSP field data, from the Bluebell
Altamont Field in the Uinta Basin, Utah in the Figure 15.54 (a) Variation of P-wave attenuation with angle from
USA (Lynn et al., 1999a). This was a tight, fractured the fracture normal, for different modelled fracture
gas reservoir in ‘sandstones’ (or lenticular sands encased radii ( frequency  30 Hz in example). (b) Variation of
P-wave attenuation with frequency and fracture radius.
in shales and carbonates), from which production
Chapman, 2002 and 2003 model, in Maultzsch et al.,
was primarily controlled by the size, orientation
2002.
and concentration of natural fractures, i.e. intersecting
joint sets.
The near-offset VSP had the source located 550 feet The time delays between the two split shear waves,
west of the well, and the three components receivers were showed a systematic variation with frequency, and of
placed at depths from 2800 feet to 8650 feet, with 50 ft course with depth. The time delay was largest (32 to
spacing. The Green River reservoir formation was 38 ms) at lowest frequencies (5–15 Hz), and smallest
located from 6687 to 8591 feet deep (approx. 2000 to (30 to 34 ms) at higher frequencies (20–40 Hz). This
2600 m). A P-wave and two orthogonal S-wave sources data was used for inverting for the theoretical fracture
were used, yielding a nine-component data set. density and fracture radius.
Shear wave splitting had long been recognised, with The roughly 2000 to 2600 m deep reservoir had Vp 
the fast shear waves giving a consistent (slightly rotating 4877 m/s, Vs  2575 m/s, density 2.60 and porosity
with depth) N46°W to N40°W (average N43°W) polar- 9.4%. For the modelling, an aspect ratio of 0.0001 was
ization. This direction was identified by the authors with chosen for the fractures (i.e. 100
m per 1 metre).
the fracture strike, but, actually as we shall see later it is Figure 15.55a shows the results of Maultzsch et al.,
closer to H max but nevertheless caused by fractures of 2003 modelling of the relative error between the meas-
large scale. Significantly, the time delay between the fast ured and the model-computed time delays. This was
and the slow shear waves showed a sharp increase with for a range of fracture densities, fracture sizes and fre-
depth at the reservoir level, indicating the presence of the quencies. The black section of the ‘bend’ in the figure
fractures. Maultzsch et al., 2003 tested the Chapman shows minimum error, for a fracture density of 0.035
model, following the Liu et al., 2003 methodology for and a fracture radius of about 3 m. Reportedly, there
interpreting the frequency dependent anisotropy at this was evidence from Lynn et al., 1995, of fracture lengths
fractured reservoir. of about 2 to 3 m, based on borehole images. The
472 Rock quality, seismic velocity, attenuation and anisotropy

Liu and co-authors emphasised the importance of


frequency on the magnitude of the S-wave anisotropy,
as demonstrated by the Chapman model. They also
emphasised that in the Bluebell-Altamont Field VSP,
the polarized fast S-waves actually showed no apparent
variation with frequency (unless at very low frequency).
However the time delays between the split shear waves
decreased as the frequency increased, due presumably
to apparent stiffening of the squirt phenomenon.
At the surface there were two joint sets, one striking
N22° to 32°W, and the other N60° to 77°E, within
about 5° of perpendicular, on average, as described
by Lynn et al., 1999. However, the maximum horizon-
tal stress (at reservoir level), was estimated as N40° to
(a)
45°W based on perpendicular borehole elongation
using four-arm calliper logs in two adjacent boreholes.
Geologically recent natural hydraulic fracturing
(gilsonite dikes) were oriented at a consistent N40° to
45°W as well.
The extensive analyses reported by Lynn et al, 1999a,
included shear wave splitting, and azimuthal P-wave
response due to the fractures (the latter was reviewed in
Chapter 14). Liu et al., 2003a concentrated on an
analysis of the frequency dependence of the S-wave
anisotropy, looking at data both above and within
the 1980 to 2590 m deep reservoir. The techniques of
rotation, band pass filtering, minimising off-diagonal
energy and so forth, are beyond the confines of our
(b) simplified treatment.
Significantly, except for the very low frequency band
Figure 15.55 (a) Relative error between VSP measurements and
Chapman model results, shows minimum at realistic
below 10 Hz, the polarizations were generally between
fracture radius a 3 m, and fracture density of 40° to 45° for the three frequency band-widths between
0.035 or about one fracture/10 m cube. (b) Per- 10 and 40 Hz, over the whole depth interval 853 to
centage anisotropy versus frequency match is excel- 2636 m, both above and within all the reservoir, (see
lent. Maultzsch et al., 2003. Figure 15.56a). This agreed well with the average N43°W
of the interpreted major stress, but did not quite agree
with the dominant jointing, as actually proposed in Liu
consistently modelled polarization angles were 43° (the et al., 2003a and Maultzsch et al., 2003.
same as the average N43°W polarization that was meas- Three distinct time-delay intervals were detected (see
ured), and the percentage anisotropy as a function of fre- Figure 15.56b and Table 15.5). Within the three depth
quency, shown in Figure 15.55b, was a reasonably good intervals there was a superimposed frequency dependence,
match; and excellent at frequencies above about 13 Hz. as shown in Figure 15.56b. The steep time delay-fre-
Liu et al., 2003a, also described methods of analysis quency data for the reservoir (interval III) is reproduced at
needed for interpreting multi-component VSP from the larger scale in Figure 15.57a, and the different gradients
fractured gas Bluebell-Altamont reservoir in the Uinta give corresponding estimates of frequency dependent
Basin in Utah. This was the subject of the above shear-wave anisotropy percentages in Figure 15.57b.
Maultzsch et al., 2003 application of the Chapman, According to Liu et al., 2003a, there were at that time
2003 dynamic model of pores, ellipsoidal microcracks very few reports of frequency–dependent anisotropy in
and aligned meso-fractures, which was used to interpret the literature, from the exploration geophysics commu-
both fracture strike and fracture dimensions at this field. nity. However, for earthquake data, Marson-Pidgeon
Shear wave splitting in fractured reservoirs and resulting from earthquakes 473

Table 15.5 Time delay-depth gradients for interpreting seismically


isotropic and anisotropic (fractured) depths at the
Bluebell-Altamont Field. Liu et al., 2003a.

Interval I, 853–1219 m Time delays linearly increase with


receiver depth increase: i.e. seismically
anisotropic, as shear wave splitting is
occurring
Interval II, 1219–2072 m Time delays almost constant: i.e.
seismically isotropic, as no further
shear wave splitting
Interval III, 2072–2636 m Time delays abruptly start increasing:
(reservoir interval) i.e. strong shear wave anisotropy of 3
to 4%, due to splitting. Attributed to
fracturing.
(a)

(b) (a)

Figure 15.56 (a) Except at low frequency (0–10 Hz), a reasonably


constant polarization at 40° to 45° is shown. (b)
Time delays show three intervals: gradient, flat,
gradient, implying anisotropy, isotropy, anisotropy.
Bluebell-Altamont Field anisotropy interpretation,
from Liu et al., 2003a, with kind provision of files
from Liu, pers.com. 2005. (see colour Plate 6).

and Savage, 1997, had reportedly shown a systematic


decrease in time delay with increasing frequency, as also
shown by the reservoir data in Figure 15.57.
Liu and co-authors suggested that if a proper mechan-
ism (or mechanisms) for this frequency dependence (b)
could be understood, as implicitly shown in the
Figure 15.57 (a) Steep time delay gradient in reservoir interval,
Chapman, 2003, dynamic model of triple-scale poros- with frequency dependence. (b) Interpretation of
ity reviewed earlier, then meso-scale fractures and their anisotropy percentage as function of frequency.
fluid flow properties could be understood. Naturally, it Bluebell-Altamont Field anisotropy interpretation,
is logical to refer to Figure 15.52 and see again from the from Liu et al., 2003a, with kind provision of files
example of Chapman triple-porosity modelling, the from Liu, pers.com. 2005. (see colour Plate 7).
474 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.59 Time-delays estimated from synthetic data for 4 fre-


quency bands (solid lines). The real data is shown
with dashed lines, showing generally excellent fit.
Liu et al., 2003b.

The estimated time delays from the Chapman model


(stepped) were finally compared with mean trends from
the reservoir (dotted lines). (Figure 15.59). The slight
discrepancy (‘a matter for future investigation’) could
perhaps, among other more complex factors, be due to
the inequality of physical aperture (E) and hydraulic
(smooth parallel-plate) aperture (e), when modelling
the aspect-ratio-dependent treatment of squirt flow
Figure 15.58 (a) The relative error between the predicted and
measured time-delay/depth, evaluated over four fre-
losses and frequency dependence in the Chapman
quency values for a range of possible fracture densi- model. In other words a given aspect ratio assumption
ties and sizes, for comparing with multi-component contains a fluid volume equivalent to thickness (E), but
shear wave VSP data acquired in the Bluebell- permeability in the ‘real world’ is governed by (e) since
Altamont field in Utah. (b) The rms error zoomed K  e2/12 if laminar. The effect of E/e  1 is caused
around the minimum, where the error is less than 5%. by joint or fracture roughness (Barton et al., 1985).
(Chapman model application, by Maultzsch et al., If looking for more reasons for partial lack of fit
2003 and Liu et al., 2003b). (see colour Plate 8). between model and one may note that Olsson and Barton,
2001, showed from coupled shear-flow experiments by
tendency for an increased relevance of larger-scale frac- Olsson, (reviewed in Chapter 16), that the E/e  1
tures or joints, for explaining shear-wave anisotropy at inequality with pure normal closure is modified in a
the lower frequencies. subtle way when there is joint shearing. There would
In a companion paper Liu et al., 2003b, also gave the then also be greater relative mobilization of shear stiff-
Maultzsch et al., 2003 prediction of fracture density ness, compared to normal stiffness, which is not of
(appropriately re-named, from crack density) and mean course modelled in the Chapman poro-elastic model,
fracture radius, obtained with the Chapman model. but could be the ‘reality’ in situ. There is also the possi-
Figure 15.58 shows the more complete result of this bility of the shear-related ‘O/R’ rotation shown in
analysis of the Bluebell-Altamont Field anisotropy, with Figure 15.2b at the beginning of this chapter.
interpreted fracture density of 0.04 and fracture radius Figure 15.60, which summarises the various azimuthal
of 3 m. These imply, from e  Na3/V, about one of relationships pertaining to the Lynn et al., 1999a
these major fractures per 9 m cube of rock. Bluebell-Altamont field investigations in Utah, in fact
Shear wave splitting in fractured reservoirs and resulting from earthquakes 475

occur down to seismic frequencies, as we have seen in


Figure 15.52.

15.9.4 The SeisRox model

There are an increasing number of numerical models


that can represent the effect of a single set of fractures
(typically vertical) on the otherwise transversely isotropic
behaviour of a layered reservoir. Two contrasting reservoir
idealizations were presented by Johansen et al., 2004 and
are shown in Figure 15.61. A reminder of the theoretical
velocity components, kindly provided by Johnsen
(pers. comm., 2005) is shown in Figure 15.62.

a) an isotropic medium without fractures


(Figure 15.61a)
b) an transversely isotropic material with one set of frac-
tures giving a horizontal symmetry axis
(Figure 15.61b)

Johansen et al., 2004, coupled a rock physics model


with a finite difference scheme for visco-elastic seis-
mic modelling (termed SeisRox), to represent the coup-
ling between fluid-filled fractures and pores. They demon-
strated a stiffening behaviour (increased velocities), as
frequency was increased. The contrasting effects of their
models of pores only (Model A) and communicating
pores and fractures (Model B), are shown in Figure 15.61.
Figure 15.60 Azimuthal fracture orientations from outcrop and The properties assumed for their fractured reservoir
core, compared to interpreted stress and interpreted layer (Table 15.6), gives insight into the level of model-
seismic data, for the Bluebell-Altamont field. Based ling detail. Note the rather large aspect ratios for the
on Lynn et al., 1995 and Lynn et al., 1999a. assumed cracks, which fall short of actual fracture (or
joint set) simulation in this particular case.
indicates a 15° discrepancy between reflection seismic To emphasise the links between stiffness, velocity
interpretation of S1, and the polarized shear-wave and attenuation, we may refer to the published article
interpretation of S1. This actual small discrepancy of the (Johansen et al., 2004) in which the vertical axes of
polarization directions for the supposed 3 m fractures, Figures 15.61a and b were given the alternative nomen-
in relation to the oriented fractures in the core (a possi- clature ‘real part of stiffness’ in place of velocity, and
ble 15° to 25° discrepancy) could be due to experimen- ‘imaginary part of stiffness’ in place of attenuation. The
tal/acquisition errors (Lynn, 2005 pers. comm.), but four sets of curves of behaviour for the six ‘V’ and ‘Q’
might also be due to a phenomenon such as illustrated components were unchanged, in this alternative nomen-
in Figure 15.2b. clature, yet in place of a velocity scale of 2.0 to 4.5 km/s,
Liu and co-authors emphasised that the Chapman they gave a stiffness scale of 5 to 50 GPa, and in place of
model reverted to a grain-size squirt flow model when an attenuation scale (1/Q) of 0 to 0.20 (or minimum
without fractures. But with the introduction of a frac- Q  5) they gave a stiffness scale of 0 to 6 GPa (the
ture set, two characteristic frequencies then existed: the ‘imaginary part of the stiffness’).
traditional lab-scale squirt frequency and a lower fre- It is of interest at this juncture, to refer to Figure
quency that depended on fracture sizes. This meant 15.33b, for the purpose of cross-discipline connections,
that dispersion (frequency dependent velocity) could suitably tempered by the dynamic situation. This diagram
476 Rock quality, seismic velocity, attenuation and anisotropy

Figure 15.61 Modelled wave speeds and attenuation as a function of frequency, using the SeisRox model, for (a) an isotropic medium with
random pores, (b) communicating pores and oriented vertical cracks in a transversely isotropic medium with horizontal sym-
metry axis. Johansen, pers. comm., 2005.

shows the empirical link between P-wave velocity from equivalent (dynamic) ‘imaginary part of stiffness’ of
refraction seismic, and the static deformation modulus M 6 GPa, following Johansen et al., 2004, would (falsely)
(as opposed to the dynamic value obviously just referred). give a low Qc estimate of about 0.2 (‘very poor rock qual-
The relationship has ‘common ground’, namely, the rock ity’) if one temporarily ignored the Edynamic  Mstatic
quality Q or Qc value, which can however be dispensed norm, and used the empirical link M  10 Qc1/3 shown
with, when the proposed direct link from seismic Vp and in Figure 15.33b.
static modulus (M) is used (Barton, 2002a).
It may be noted, for purpose of comparison, that the
above 2.0 to 4.5 km/s range of velocities, with their equiv- 15.9.5 Numerical modelling of
alent (dynamic) ‘real part of stiffness’ of 5 to 50 GPa, have dynamic joint stiffness effects
a predicted static moduli (M) range of 3 GPa to 22 GPa,
respectively. The maximum attenuation (1/Q) scale value In the foregoing double-porosity models we have wit-
of 0.20, giving the particularly low Qseis value of 5, and an nessed the use of crack representation by aspect ratios.
Shear wave splitting in fractured reservoirs and resulting from earthquakes 477

(a)

(b)

Figure 15.62 The stiffness matrices showing the velocity components modelled in Figure 15.61. Johansen, pers. comm., 2005.

Important insight into the influence of (dynamic) joint cross-hole seismic at frequencies of several kHz. Monsen
stiffness and joint spacing effects on seismic dispersion is first explored the influence of single fractures, and assumed
provided by the modelling of Monsen, 2001. The motiv- dynamic normal stiffness values (100, 1,000 and
ation for his study was the interpretation of high fre- 10,000 MPa/mm) equivalent to widely different normal
quency investigations of rock masses, in which the wave stress levels, from e.g. close to surface to kilometre depths,
length may be less than the joint spacing, for instance in as this range indicates (see Chapter 16).
478 Rock quality, seismic velocity, attenuation and anisotropy

Table 15.6 Parameters used to compute effective visco-elastic


properties of reservoir rock. (Johansen et al., 2004).

Porosity (%) 15
Permeability (mDa) 50
Clay content (%) 15
Viscosity (cP) 1
Fluid density (kg/m3) 1000
Fluid velocity (m/s) 1500
Aspect ratio quartz-related pores 0.15
Aspect ratio clay-related pores 0.027
Aspect ratio cracks 0.01
Crack density 0.1
Effective density (kg/m3) 2383
(a)

The soft-fracture near-surface simulation, showed the


single fracture to be a good reflector, with a wide range
of low to medium transmission coefficients (T) until fre-
quencies were lower than 1 kHz, when T became unity.
The stiffest, deepest simulation, showed transmission
coefficients of mostly 1.0, except at frequencies higher
than 100 kHz. In effect the time delay here was close to
zero except at the highest frequencies. There was natu-
rally more delay and attenuation as the dynamic frac-
ture normal stiffness was reduced.
The effect of multiple parallel fractures on the group
(b)
velocity, as compared to a ‘sedimentary rock’ matrix
velocity of just below 2490 m/s is illustrated in
Figure 15.63. The same range of dynamic normal stiff- Figure 15.63 (a) Group velocity versus frequency for three frac-
ture spacings, at a constant (near-surface, and very
ness as above was used (i.e. from near-surface to several
low) dynamic normal fracture stiffness of
kilometres equivalent depths), and three alternative
100 MPa/mm. This behaviour corresponds quite
spacings of 0.1, 1.0 and 10 m were used. The reduced accurately with near-surface rock quality (Q) versus
velocity with small fracture spacing is of course a famil- seismic velocity expectations. (b) Group velocity ver-
iar trend in near-surface rock mass quality (Q) relations sus frequency with a constant fracture spacing of
to Vp, with limited effect on velocity (and rock quality), only 0.1 m, and the full range of dynamic normal
when the spacing is as high as 1 or particularly 10 stiffnesses of 100, 1,000 and 10,000 MPa/mm. The
meters. benefit of cross-well, high frequency surveys is
However, as also suggested by the results in demonstrated quite well by these scoping studies.
Figure 15.63, when equivalent depth is of kilometre to After Monsen, 2001.
several kilometres magnitude, a fracture spacing as small
as 0.1 m has small, or negligible effect on the perceived
velocities with dynamic stiffnesses of Kn  1,000 and UDEC, the provision of realistic input data concerning
10,000 MPa/mm, unless at frequencies beyond 10 kHz. joint or fracture dynamic stiffnesses is fundamental for
This is a good illustration of the benefit of high fre- a representative result. In Figure 15.64, we see the effect
quency cross-well measurements, in particular in view of a more realistic ‘jointed rock’ (in 2D), on the attenu-
of the long transmission distances of high frequency ation of wave amplitude, on velocity reduction, and on
waves at depth, when joints and fractures are stiff, as frequency change. The centre of each model (intact or
also referred to in Chapter 14. jointed) is the source location, with a measurement
When modelling wave propagation through jointed point at the boundary in each case. (Non-reflecting/
or fractured rock with distinct element codes like absorbing boundaries are of course used). In the cases
Shear wave splitting in fractured reservoirs and resulting from earthquakes 479

Figure 15.64 An illustration of wave propagation differences between intact and realistically jointed (but 2D) models, with dynamic nor-
mal and shear stiffness input estimates appropriate to the dynamic wave form. Note the multiple effects of (vertical) wave
transmission through the layered and jointed medium: i.e. the delay, the amplitude reduction (i.e. attenuation), the velocity
reduction and the frequency reduction. (Monsen, 2005 priv. comm).

illustrated, the dynamic stiffnesses assumed were within Snow, 1968, and is utilised for grouting predictions in
the range studied by Monsen, 2001. There is at present rock engineering (e.g. Figure 15.49).
no matrix flow modelling in UDEC. Brown et al., 2002 assumed that both the fracture
permeability tensor and the fracture compliance tensor
would ‘diagonalise’ in the same principal coordinate
15.9.6 A ‘sugar cube’ model system. The model predictions for P-wave and S-wave
representation phase velocities, contrasting the matrix velocity with
the effect of either a drained or brine-filled fracture net-
A practical ‘sugar-cube’ model with three orthogonal work is shown in Figure 15.65.
joint or fracture sets was proposed by Brown et al., 2002, The authors noted that when the compliance of the
as a means of integrating seismic data to production data. fractures was large, as in the field, where fractures tend
The authors combined the ideas of Oda, Kachanov, to be weaker, the effect of fluid changes on the velocity
Schoenberg and Sayers to produce the rather obvious rep- would tend to be dominated by effective stiffness
resentation of a multiply-jointed rock mass, where three, increases due to e.g. brine saturation compared to
more or less mutually perpendicular joint sets is almost a drained or gas filled. The authors drew parallels to the
rule rather than an exception. Such a representation for Natih field experiences described by van der Kolk et al.,
the permeability of rock masses at dam sites was used by 2001, which were reviewed earlier in this chapter.
480 Rock quality, seismic velocity, attenuation and anisotropy

15.10 A porous and fractured


physical model as a numerical
model validation

The technique for constructing synthetic, fractured


sandstones by Rathore et al., 1995, was referred to earl-
ier, as a means of basic validation of numerical models.
The (future) cracks are placed as metallic discs in a porous
sand-epoxy mixture, and subsequently leached out,
leaving empty voids. They thus have a known aspect
ratio, size, position and orientation and are a close
approximation to basic aspects of dual-porosity theor-
etical models. (Both however have the potential weak-
ness in relation to reality, that stress and displacement is
not transferred across the crack faces, so E  e, since
roughness may be absent).
The crack geometry studied by Skjærstein and Fjær,
2000, consisted of 1425 circular parallel cracks per dm3,
with all cracks of 8.25 mm diameter and 20
m aperture.
The crack density (number of cracks per unit volume 
crack radius cubed) was 0.1. A representation of the
crack-modelling principles is shown in Figure 15.66a.
The authors pointed out that the (Rayleigh wave)
scattering was strongly dependent on the wavelength-
to-crack size ratio, and therefore on the frequency for a
given crack size. Since the velocity anisotropy was
not equally dependent on frequency at high ratios of
wavelength-to-crack-size, the big discrepancy between
Vmin/Vmax and Amin/Amax should decrease with decreas-
ing frequency.
With the crack size studied by Skjærstein and Fjær,
2000, the P-wave velocity anisotropy was greatly reduced
by saturation, while in earlier models studied by
Rathmore et al., 1995, there was significantly less reduc-
tion in velocity anisotropy with saturation, due prob-
ably to the smaller size of cracks that were used, giving
different conditions of pore pressure equilibration.
Skjærstein and Fjær, 2000, testing first dry then sat-
urated cracks, found that the attenuation was strongly
dependent on the incidence angle, as indicated in Figure
15.66b and in Figure 15.67. The attenuation, which
was caused by scattering, distorted the wave forms and
changed their frequency content. The angular depend-
ence of attenuation was affected by the frequency (e.g.
100, 250 or 500 kHz). In the case of the 100 kHz
excitation pulse applied across the 16-sided prismatic,
fracture-filled sample, the wavelength-to-crack diame-
Figure 15.65 A ‘sugar-cube’ model of compliance and permeabil- ter ratio was in the range of 1.7 to 3.0.
ity, for integrating seismic and flow measurements. Under the experimental conditions investigated by
Brown et al., 2002. Skjærstein and Fjær, 2000, the attenuation anisotropy was
Shear wave splitting in fractured reservoirs and resulting from earthquakes 481

Figure 15.67 (a) Recorded P-wave signals propagating parallel or


normal to dry cracks in a synthetic, crack-bearing
porous sandstone (D/t  8.25/0.02 mm/mm) from
Fjær, 1998. (b) Recorded P-wave signals when using
three different excitation frequencies, in a dry crack-
Figure 15.66 (a) The prismatic, sixteen-sided, crack-bearing syn- bearing synthetic, porous sandstone. A five-cycle
thetic sandstone used by Skjærstein and Fjær, 2000. pulse was used at 500 kHz. These samples had 1425
(b) A comparison of P-wave and polarized S-wave circular parallel cracks/dm3 with the same D/t meas-
(qS1 and qS2) results using variable angle of inci- urement as above. Skjærstein and Fjær, 2000
dence, for the two cases of i) through-going spot-
welded fractures: open symbols, and ii) small parallel
cracks: closed symbols. Squares are P-waves,
circles/triangles are S-waves polarized parallel/quasi- closed symbols in this figure represent 5.5 mm diameter
normal to the cracks. Fjær, 1997. cracks of 20
m aperture, while the open symbols repre-
sent the through-going but spot-welded cracks, where the
contact spot diameters were 1.4 mm and the open parts
much more visible than the velocity anisotropy. Also at the same 20
m. Each case studied was dry. These two
higher frequencies, where the P-wave velocity anisotropy special (and rather theoretical) geometries were chosen to
vanished, the attenuation anisotropy remained high represent a Hudson, 1981, type of crack distribution, and
and clearly visible. Likewise, when saturated, causing a a White, 1983, model with spot-contact points on the
significant reduction in the P-wave velocity anisotropy, larger fracture planes.
the attenuation anisotropy remained high. The authors When the wave-length was many times larger than
recorded Vmin/Vmax  0.80, and Amin/Amax  0.01 for the crack diameters, the cracked material effectively
the same wave frequency. appeared as a homogeneous medium with reduced
In Figure 15.66b, we may note one of the ‘complicat- stiffness, while at higher frequencies, the smaller wave
ing’ factors concerning the diagnostic use of P- and S- length-to-crack diameter ratio effectively made the
wave polarization. Fjær, 1997, presented the P-wave and medium appear more heterogeneous. Shear wave split-
polarized S-wave results in parallel for i) through-going ting with fast and slow components then showed
but spot-welded cracks, and ii) small parallel cracks. The greater contrast.
16 Joint stiffness and compliance
and the joint shearing
mechanism

This final chapter is designed to act as a cross-discipline set of ‘perfectly’ aligned, or partially aligned, sets of cracks,
reference point between rock mechanics and engineering fractures (or rock joint analogues) was seen. The mathe-
geological behaviour in the ‘static’ world of slow-and- matical complexity of multiple sets clearly presents numer-
macro deformation processes, and the geophysicists ical problems, but no doubt such will be solved in due
‘dynamic’ world of fast-and-micro deformation and atten- time, as it was, with simpler boundary conditions, in rock
uation processes. That there are important links between mechanics, due to the need for modelling hydraulically
the two in terms of joint or fracture compliance and its communicating, deforming, multiple joint sets, but with-
inversion: stiffness, and in terms of rock quality, deform- out the dynamic interaction with grain-size pore space.
ation modulus, and seismic quality, has been established The deliberate addition in Figure 16.1, of a secondary
in various contexts in the chapters of Part II. In particular, ‘crack set’ to the Tod, 2002 conceptual model of a variably
this last chapter attempts to extend current thinking oriented single set of cracks, fractures or joints, is to
regarding fractured reservoirs, ‘open’ joints, and assumed emphasise the obvious: that fractured (or jointed) reser-
H max parallelism, to also embrace the possibility, even voirs often have at least two sets of joints or fractures,
probability, that multiple-joint-sets and shear-dilation- that can assist in the drainage of the matrix towards the
conductivity coupling are needed, due to the inevitable wells. Primary drainage pathways are usually considered
tendency for joints in less competent rocks to close at
reservoir depths, even when under the influence of ‘only’
h min. Shear wave splitting and polarization from two sets
of conjugate (or H max straddling) joints can also appear
to satisfy ‘open joints were parallel to H max’ assumptions.
The important improvement would be that rock mechan-
ics theory is not violated, when assuming that unsheared
joints in weaker reservoir rocks can be conductors, when
actually under high levels of effective stress. Shearing to
great depth has been verified in other areas of the earth-
sciences, and the need for adoption in the geophysical
interpretation of petroleum reservoirs is discussed, with
simple joint-model illustrations. The ‘critical shearing
crust’ is interpreted here in terms of the non-linear Barton-
Bandis shear strength and coupled behaviour constitutive
model. An evaluation of the importance of joint rough-
ness at several scales, and the need for dilation-corrected
stress transformation in the earth sciences is also treated.

Figure 16.1 A secondary set of joints (i.e. natural fractures) have


16.1 Some important non-linear joint been added to the Tod, 2002 schematic, to emphasise
and fracture behaviour modes that in many cases there will be additional intersect-
ing sets of joints in fractured reservoirs. Joints may
In the Chapter 15 review of some of the recent dynamic, also be under a state of shear stress, rather than the
poroelastic, multiple-porosity modelling, mostly only one implied normal-stress-only.
484 Rock quality, seismic velocity, attenuation and anisotropy

to be along the set that approximates the H maximum Numerical discrete element modelling of these three
direction, as we have seen, with some variations, in the categories of jointed structures, was performed by
shear wave splitting analyses reviewed in the last chapter. Chryssanthakis et al., 1991, with UDEC-BB, incorp-
This conventional assumption may be modified how- orating the Barton-Bandis non-linear and scale-depend-
ever, if the rock is weak and porous, since conducting ent joint behaviour laws, to confirm this type of rock
joints might be too tightly closed, or not exist, at depths mass stress-deformation behaviour. Type A and Type C
of kilometres, unless there was shearing. This aspect has can be considered typical of many reservoir jointing
dominated the selection of material for this chapter. scenarios (greatly simplified), with sedimentary bed-
In case of significant non-alignment of the present-day limited jointing in Type A, often with a vertical major
H maximum direction with the dominant set (or sets) of principal stress, but with an intermediate horizontal
joints, then shear-dilation-conductivity coupling becomes principal stress that could cause fracture-set alignment.
a potential mechanism for effective reservoir drainage Type C, representing conjugate jointing, is of tectonic
from matrix-to-joints-to-producing well. The range of origin, as for instance in the anticlinal Ekofisk reservoir
effective block sizes illustrated in Figure 14.1 (1, 2 and 3 in jointed chalk. The deformation, joint-shear and prin-
are examples) will then also come into consideration, cipal-stress plots from this numerical modelling are
since block size will determine how much shear-dilation- shown in Figure 16.3.
permeability coupling has been possible, especially with The one-dimensional application of the Barton-Bandis
elevated pore pressure. normal closure modelling, shown in Figure 16.4, from
Scaling of the small-scale roughness JRCo to effective Barton et al., 1985, follows the multiple load-unload con-
roughness JRCn for the appropriate block size, follow- solidation cycles performed on numerous natural rock
ing the Barton-Bandis constitutive model, suggests that joints by Bandis, 1980. In the discrete element UDEC-
case 3 with the largest block size, can have presented the BB modelling performed in rock mechanics analyses of
lowest shear stiffness (and shear strength) to the historic e.g. slopes, tunnels and caverns for civil engineering use,
anisotropic stress that may have caused shearing in the we automatically follow the ‘fourth cycle’ which is
past. A sheared-joint situation would presumably tend assumed to approximate the ‘undisturbed’ in situ condi-
to enhance the velocity anisotropy (both for azimuthal tion. There is naturally an exaggerated hysteresis (and
P-waves, and for split shear-waves). closure) in the first cycle of a stress-closure test on a sam-
A cross-disciplinary reference to the (macro- pled ( disturbed) joint. In the context of the micro-
deformation) hyperbolic normal-closure behaviour of displacements experienced by joints with the passage of
rock joints, and to the scale effect on shear strength caused seismic waves, it is clearly the consolidated state one is
by relative block-size, are key items in this ‘sheared-earth’ concerned with.
thinking. These macro-deformation aspects of jointed Hatchell and Bourne 2005, have recently highlighted
rock behaviour are shown at the top of Figure 16.2, taken the importance of hysteresis in the interpretation of time-
from Bandis et al., 1983, and 1981. This macro-deforma- lapse time shifts surrounding depleting, compacting
tion behaviour is shown here, in view of the influence it reservoirs. Depletion induced time shifts have opposite
could have on the micro-deformation compliances of joints sign inside and outside the reservoir. Laboratory testing of
that were ‘lying close’ to the respective rock mass stiffness velocity-effective pressure needs to also be focussed on
curves. Of course their will be differences between the unloading and on strain behaviour.
inverted magnitudes of dynamic compliance and ‘static’ Models that use linear slip theory, following earlier
stiffness, but differences may be minor when rock quality Schoenberg work, and therefore require fracture com-
is reasonably high. pliances rather than aspect ratios, are gaining ground in
The assumed stress-deformation behaviour of three cat- geophysics, as we have seen from some of the modelling
egories of rock mass, as visualised by Barton, 1986, are reviewed in Chapter 15. Here we will consider a remark
shown below the normal-closure (N) and shear- by Gurevich, 2002, who describes the theoretical con-
displacement (S) components, in Figure 16.2c, and in struction of a poro-elastic model with aligned fractures
Table 16.1. There is large scale, up to 8 m3 biaxial flat-jack that obey linear-slip theory, and are characterized by
loading, and uniaxial plate-loading verification of these ‘excess’ normal and tangential fracture compliances,
three categories of non-linear and hysteretic behaviour. following Schoenberg and Sayers, 1995. Gurevich com-
(The linear load-deformation line in case B appears to be mented on the fact that several authors had expressed,
due to two partly opposed non-linear mechanisms). intuitively, the likelihood of a link between normal and
Joint stiffness and compliance and the joint shearing mechanism 485

Figure 16.2 a) Non-linear, hysteretic normal-closure behaviour of rock joints. b) Shear strength dependency on block size, also showing
strong non-linearity and two reasons for strong shear stiffness scale effects. c) Conceptual pressure-deformation curves due to N
and S components, or their combination. From top left: Bandis et al., 1983, Bandis et al., 1981, and Barton, 1986.

Table 16.1 Three characteristic load-deformation behaviour-modes for rock masses.

Type Dominant mode Shape Hysteresis Lateral expansion* Poisson 

A normal concave small small low


B normal  shear linear moderate moderate medium
C shear convex large large high

* Under conditions of one-dimensional strain the relative magnitudes of these differences in behaviour will be
reduced. Reduction in pore pressure, and compaction of the matrix, would allow a Type C mechanism
in a compacting reservoir to develop further, as for instance in the Ekofisk reservoir.

tangential fracture compliances and the porosity of the implied linkage to porosity, since the term JCS, the
host medium. joint wall compressive strength, and its ratio with effec-
From the rock mechanics world of inverted compli- tive normal stress (JCS/ n) is an important component
ance, i.e. stiffness (where Kn and Ks stand for normal of the equation for shear stiffness and peak shear strength
and shear stiffness in units of MPa/mm), there is also an (Barton and Choubey, 1977), and of the hyperbolic
486 Rock quality, seismic velocity, attenuation and anisotropy

equation for normal stiffness (Bandis, 1980). The value been mineralization at some time, as shown in the
of JCS is obtained from rock mechanics index testing, photographs (Figure 16.5, a and b) of a wave-cut plat-
using a Schmidt (L-) impact hammer, the rebound of form in dolomite, from Kimmeridge Bay in Southern
which (10 to 95% approx.) can be linked to uniaxial England. Note the complete cementation of all joint sets
compression strength c (or UCS), also using rock in Figure 16.5c, which probably removes the need for a
density, following Miller, 1965. Clearly porosity will roughness and wall-strength based classification method.
therefore affect this conversion of rebound to JCS
(or to matrix UCS). Hence the connection to stiffness,
and therefore perhaps compliance. 16.2 Aspects of fluid flow in
The Schmidt hammer is used as this is almost deforming rock joints
the only way to record the reduced strength of the
few millimetres of (eventually weathered, or altered) Several recent workshops have focussed strongly on seis-
joint walls. In other words JCS  c, unless there has mic anisotropy, suggesting an active recognition of the

Figure 16.3 Distinct element modelling with UDEC-BB code, with the Barton-Bandis joint behaviour sub-routine, demonstrates the rela-
tive magnitudes of the shear and normal deformation components. Top: deformation, Centre: joint shearing, Bottom: princi-
pal stresses. Chryssanthakis et al., 1991.
Joint stiffness and compliance and the joint shearing mechanism 487

over-riding importance of dominant fracture sets (or (Makurat et al., 1990), and first used for extensive studies
joint sets in engineering geology terms) for drainage of of the conjugate joints (or fractures) in the Ekofisk chalk,
reservoirs. This is particularly important where matrix to investigate oil and brine flow under normal and shear
permeabilities are low, yet storage in the form of porosity, deformation, also with the effect of elevated temperature.
obviously high enough for commercial development, Various aspects of this work with laboratory-scale testing
sometimes to an exceptional degree. The Ekofisk field in of reservoir joints and reservoir-scale deformation mod-
the North Sea, is expected to have at least 80 years of pro- elling in the chalk were described by Barton et al., 1986.
duction from this well jointed reservoir. Although dispersive (frequency-dependent) attenu-
ation and velocity are dynamic concepts, involving micro-
responses in the interior of any rock joint, it seems
16.2.1 Coupled stress-flow intuitively likely that the macro-responses of joints to
behaviour under normal coupled deformation and flow can represent
closure ‘end-members’ for the dynamic loading and squirt
attenuation concepts of geophysicists.
Coupled stress-flow-deformation joint behaviour has The normal and shear compliances (units of m.GPa1)
long been of concern in rock mechanics, where there is presently utilised in geophysics, have their rock mechan-
a ‘standard’ test called the coupled shear flow test ics inverse values, termed normal and shear stiffness (units
(CSFT), developed by Makurat at NGI, in Norway, of GPa/m). Their (inverted) magnitudes are remarkably
similar, especially in the case of Kn, suggesting some yet
to be determined relationship, albeit perhaps in an
‘end-member’ role.
Under normal closure (i.e. the perpendicular incident
wave case), the hydraulic aperture (e) of a rock joint, as
measured in a (macro) flow test, is related to the larger
(mean) physical aperture (E) by the roughness-related
empirical equation (Barton et al., 1985):

E2
e ≈ (16.1)
JRC 2.5

(where E  e. The joint roughness coefficient JRC,


varies from 1.0 for smooth joints: giving, implicitly e  E,
to about 20 for extremely rough tension fractures).
There seems every reason for these ‘macro-roughness’
and ‘macro-flow’ concepts to apply as boundary condi-
tions, in the case of intrinsic attenuation, at least in
broad terms. The fundamental importance of joint or
fracture (macro) roughness is demonstrated by empir-
ical data in Figure 16.6, and Figure 16.7 shows the form
of this empirical relation.
An illustrative example of the physical mismatch of
E and e, and the practical effect this can have on e.g. the
groutability of a near-surface rock mass is illustrated in
Figure 16.4 Normal closure and conductivity cycles and hysteresis
from the BB model. An imaginary injection-produc-
Figure 16.8. This ‘cubic’ model of Snow, 1968, with
tion cycle is highlighted in gray. Barton et al., 1985. additions by Barton, 1986 is a convenient vehicle for
(The importance of hysteresis in time-lapse interpreta- planning the choice of suitable grout particle sizes (d95),
tion both inside and outside a compacting reservoir with the rule-of-thumb E  4 d95. When there is insigni-
was recently emphasised by Hatchell and Bourne, ficant mismatch of these quantities E and d95 due to high
2005.) stress or due to the low permeability of tight joints, then
488 Rock quality, seismic velocity, attenuation and anisotropy

higher (locally deforming) injection pressures are used


e.g. Barton, 2003. Note the interpreted increase of the
spacing of water-conducting joints or fractures with depth,
based on a statistical analysis of the Lugeon water injection
test, which was depicted diagramatically in Figure 15.49,
in Chapter 15.
The above very simple single-parameter approach to
characterizing the mismatch of the mechanical or phys-
ical aperture (E) and the usually smaller hydraulic aper-
ture (E), does not seem to be known in the geophysics
literature reviewed. There is of course data from an
increasing number of research projects that has supported,
(a)
and continues to support, the fundamental mismatch
of these quantities, with early sources of such data listed
in Figure 16.6.
Key experimental research in this area includes Wither-
spoon et al., 1979, Barton et al., 1985, Gale, 1987, Gen-
tier 1986, Pyrak-Nolte et al., 1990, Makurat et al., 1990,
and Cook, 1992. A more recent review by Renshaw,
1995, and aperture modelling contributions from
Sisavath et al., 2003 and Liu, 2005 emphasise the interest
in this ‘detail’, which in fact is fundamental in controlling
the world’s supply of oil, gas and water. Liu also consid-
ered that naturally fractured formations under reservoir
conditions would have very small apertures (0.1–10
m),
very rough fracture surfaces, and a large proportion of the
(b) fracture surfaces in contact with each other.
This opinion, which is shared if only normal closure
(i.e. interlock) is considered, turns out to be the least
productive condition, exceeded only by more planar joints
in weaker rock, which would hardly conduct fluids if
tightly closed. With somewhat lesser roughness and slight
shear, a fundamental enhancement to these (too) limited
apertures is possible, as we shall see below.

16.2.2 Coupled stress-flow


behaviour under shear
deformation
(c)
When there is shear displacement (i.e. a discontinuity in
Figure 16.5 A wave-cut platform in a jointed dolomite bed. These the shear stress-load behaviour), experiments show that
beds occur at intervals in the Kimmeridge shale, out- the roughness ‘iso-curves’ of Figure 16.7, are cut nearly
cropping in Kimmeridge Bay, Dorset, England. The
at right angles in the E/e versus e space. Some of these
joints show a) implied JCS  c due to weathering and
trends are seen in Figure 16.6, from a slightly sheared,
preferential wave erosion, and b) implied (local)
JCS  c, due to subsequent mineralization of domi-
heated block test (see ‘N–S’, ‘E–W’ etc.). Esaki et al.
nant conducting joints. c) A fine example of joint 1995, confirmed these ‘perpendicular’ trends, and Olsson
cementation which may prevent the use of normal joint and Barton, 2001, formalised the results, following further
characterization techniques. Widemouth Bay, near CSFT testing by Olsson (Figure 16.9) for the case of
Bude, Devon, England. (See Plate 9). the mismatch of (e) and (E) that occurs when there is shear
Joint stiffness and compliance and the joint shearing mechanism 489

Figure 16.6 Mismatch of (e) and (E) due to joint or fracture roughness. Barton and Quadros, 1997, (updated from Barton et al., 1985)

Figure 16.7 An empirical model for the mismatch of E and e, from Barton et al., 1985.

(and some gouge production). The following empirical The JRCmobilized concept of joint or fracture shear strength
relation was found to apply: mobilization is explained with a low stress example in
1 (16.2) Figure 16.10. The early mobilization of friction, fol-
e  E 2  JRCmob lowed later by dilation and roughness mobilization and
490 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.9 Data points are from CSFT (coupled shear-flow tests)
by Olsson, in Olsson and Barton, 2001, showing ‘per-
pendicular’ trends in relation to the E-e-JRC model, due
to shearing and gouge formation. The ‘sloping-to-the-
right’ curves are relevant to shear behaviour, and are sim-
ilar to trends shown by Esaki et al., 1995. In other words
shearing, dilation, and possible gouge formation causes
data to climb the curves (for JRCmob), shown to the left.
Figure 16.8 A cubic-network approach to near-surface permeability
at dam sites (after Snow, 1968), with the addition of
aperture differentiation and roughness discrimination
by Barton, 1986.
clearly be affected by minor (or larger) amounts of (pre-)
shearing. Figure 16.11 shows a set of experimental data
(solid black points), from a self-weight, very low stress
final roughness degradation, was quantified by Barton, ‘CSFT’ from Maini and Hocking, 1977. The ‘twin’
1982, and forms an important basis for the shear-part of curves in the other three diagrams were obtained using
the Barton-Bandis constitutive model. the shear, dilation and flow coupling given by the Barton-
Note the location of ‘dilation begins’, corresponding to Bandis model, based on the JRCmob concept shown in
the initial mobilization of roughness. In the ‘roughness Figure 16.10. Note the large change of measured and
destroyed’ post-peak area there is a physical reduction in modelled permeability due to slight dilation, for the case
JRC, in other words retesting would show a reduced of these very planar (low JRC) cleavage joints.
peak strength for the joint in question. The early mobil- A more recent coupled shear-flow test performed at
ization of friction (prior to dilation), can be envisaged as higher stress by Olsson, using controlled normal stiffness
a potential contributor to the intrinsic frictional attenu- loading in a large direct shear box apparatus (see Olsson
ation mechanism, but would presumably involve and Barton, 2001), is shown in Figure 16.12. The term
(almost) recoverable micro-slip. ‘krm’ in the individual figures stands for applied normal
The other part of intrinsic attenuation involving stiffness, with ‘krm  0 kN/mm’ representing a conven-
fluid flow: potential viscous shearing, or ‘squirt’, will tional, constant normal load, zero-stiffness test.
Joint stiffness and compliance and the joint shearing mechanism 491

Figure 16.10 The JRCmobilized concept of Barton, 1982. This shows


shear strength development for rock joints expressed
as a dimensionless JRCmob/JRCpeak, and reduction of
this ratio with dimensionless (/peak) displacement.
Peak shear strength is given by ␶  n tan [JRC log
(JCS/ n)  r], from Barton and Choubey, 1977.
This diagram also gives a direct method for predicting
the theoretical (maximum) permeability development
with displacement and dilation. For this one uses the
JRC-based E  e inequality, and the initial-aperture-
enhancing dilation. (For gouge-reduced permeability
estimation, one uses the JRCmobilized relation, as shown
in Figure 16.9).

A CSFT test by Makurat (Makurat et al., 1990) is Figure 16.11 An early coupled shear flow test by Maini (in Maini
shown in Figure 16.13. This was selected as it shows that and Hocking, 1977), and an attempt to match this
with gouge production caused by shearing damage, there by estimating the shear-induced increase in joint
will be a discrepancy between the dilation-based perme- aperture due to full application of the dilation effect,
ability modelling (assuming full application of the dila- using the Barton-Bandis model. (Barton, 1982).
tion-enhanced aperture), and the experimental result. The
measured permeability increase with shearing was less than
expected from the measured dilation, due to gouge pro-
duction compromising the effect of the increased aperture. 2 mm of shear) but with reduced rate at larger shear
In two of the above cases, the theoretical effect of the deformation.
modelled dilation on permeability is shown. In the low From the point of view of attenuation potential in
stress case (Figure 16.11), a reasonably good match to the the reservoir situation, the presence of ‘open’ fractures
dramatic, two-orders-of-magnitude increase in per- caused by limited shear, may be an important additional
meability with less than 0.6 mm of shear is suggested. The effect for dynamically induced (micro) fluid flow, and
tests by Olsson (Figure 16.12) show a smaller, but signif- squirt losses, during the passage of a wave front. Sheared
icant increase (e.g. one order of magnitude increase with fractures may represent some of the ‘open’ fractures
492 Rock quality, seismic velocity, attenuation and anisotropy

16.3 Some important details


concerning rock joint
stiffnesses Kn and Ks

Normal stiffness Kn is defined as the normal stress


increment required for a small closure of a (usually very
tight) joint or fracture, at a given level of effective stress.
It is therefore clear that it is usually of significantly larger
value than the shear stiffness Ks which is the stiffness in
shear, usually taken as the average slope up to the shear
strength-displacement peak.
Since compliance in both normal and shear directions
Figure 16.12 CSFT tests by Olsson using controlled normal stiff-
is becoming of such importance in recent geophysics
ness during the shearing. One order of magnitude
increase in permeability with 2 mm of shear is indi-
modelling, some of the parallel, but macro-displacement
cated. (Olsson and Barton, 2001). aspects of the inversed parameters, Kn and Ks will be
further explored in this section.
It is believed that when rock quality is good, and
joints are hard, there will be great similarity in magni-
tude between Kn and its dynamic ‘inverse’ 1/ZN, where
ZN is the dynamic normal compliance. This is probably
because of the similarity, in good rock conditions, between
E dynamic and E static, as investigated in Part I. Between Ks
and its dynamic ‘inverse’ 1/ZT, where ZT is the dynamic
shear compliance, there is likely to be a significant dif-
ference, due to the lesser likelihood that block size will be
the determinant factor, as it is for the magnitude of the
pseudo-static shear stiffness Ks.
When rock is of poor quality, and joints are altered or
clay-bearing, there will be greater discrepancies between
each of the above three pairs of parameters, but Kn and
1/ZN should remain within an order of magnitude of
each other. These aspects have been discussed in slightly
different context, in a number of chapters in Part II.
Barton, 1972, Barton and Hansteen, 1979, Bandis,
1980, Bandis et al., 1981 and Barton, 1982 have each
emphasised the scale-dependent nature of shear stiffness
Figure 16.13 CSFT test by Makurat, showing the experimentally
Ks. An assembly of various sets of data is reproduced in
measured increase in permeability during 2.5 mm of
Figure 16.14, from Barton, 1982.
shearing on a joint in weathered gneiss, compared to
the modelled, much larger permeability increase, if
By means of characterizing joint samples recovered in
one assumes no blockage to flow by gouge produc- core, using the JRC (joint roughness coefficient) and JCS
tion. Makurat et al., 1990b and Bandis et al., 1985. (joint wall compressive strength) parameters, it is possible
with Barton-Bandis joint modelling to simply estimate
the above scale-dependent shear stiffnesses. An example,
often referred to, and they could be an important using two different laboratory (core) sized samples, is
source of in situ anisotropy, as described in Chapters 14 shown in Figure 16.15 for an assumed block-size range
and 15. However, with higher effective normal stresses, of 100 mm to 10 m. The three diagonal pairs of curves,
and with the likely presence of some gouge, making for represent shear stiffnesses at three effective normal stress
a tortuous flow path and generally tighter apertures (i.e. levels. Techniques for estimating the joint characteristics
smaller aspect ratios, despite shearing), flow resistance is JRC and JCS, developed by Barton and Choubey, 1977,
likely to be significant. are illustrated in graphical format later in this chapter,
Joint stiffness and compliance and the joint shearing mechanism 493

Figure 16.15 Estimation of peak shear stiffness based on extrapo-


lation of two different joint characters (as observed
at core-scale) using Barton-Bandis scaling, from
Barton, 1982. The extent of direct relevance of these
‘macro-displacement’ stiffnesses to seismic ‘micro-
displacements’, using the stiffness-compliance inver-
sion, with adjustment for the Edynamic # ,static
Figure 16.14 Experimental ‘large strain’ data for the shear stiffness inequality, is presently more uncertain, when moving
(Ks) of rock joints, clay-bearing discontinuities, and in the direction of larger block sizes.
model tension fractures. A very strong scale depend-
ence, and a strong (effective normal) stress dependence ‘sample-dependent’, i.e. if a larger, more continuous,
are each implied. These are fundamental properties of water-bearing, feature is sampled, it will likely have
rock masses in ‘macro-deformation’ processes. Barton, lower normal (and shear) stiffness than a less continuous,
1982. probably unweathered and usually rougher joint surface
(Barton, 1990b).
when exploring the effects of stress on modelled joint As shown in extensive studies by Bandis, 1980, and
apertures. Bandis et al., 1983, the normal stiffness of joints Kn is
Normal stiffness (Kn), by comparison to Ks, may especially sensitive to sample disturbance (i.e. inevitable
not be scale-dependent, although it is undoubtedly unloading when recovering samples), and the first load
494 Rock quality, seismic velocity, attenuation and anisotropy

cycle as depicted in Figure 16.16, shows much lower It is clear from inspection of data in Figure 16.17,
initial stiffness and much larger hysteresis. Lower nor- and elsewhere, that these average third cycle, initial
mal stiffness is also seen following shearing, due to a normal stiffness values of roughly 30 to 120 MPa/mm,
certain mismatch of the previously interlocking rough- correspond to effective normal stresses in the lowest
ness. The reduced contact area causes the reduced stiff-
ness, with typically 1⁄2 to 1⁄8 reduction, depending on stress
level (Bandis et al. 1983). Adverse effects on the dynamic
compliances ZN and ZT are likely too.
The contrasting normal stiffnesses from first, second
and third cycles of loading for fresh joints in hard rock are
shown in Figure 16.17. Corresponding curves for joints
in weaker rocks or weathered rocks are markedly less
steep, and show much more hysteresis on the first cycle.
The relative effects of different rock types on normal
stiffness including the weaker sandstones, siltstones
and limestones are illustrated in Figure 16.18, from
Bandis, 1980. Here we see the total (rock  joint)
deformation Vt.

16.3.1 Initial normal stiffness


measured at low stress

A major tabulation of initial normal stiffness (Kni), nor-


mal stiffness and maximum closure ranges for first, second
and third load cycles for numerous rock joint samples
with widely different JRC and JCS values was assembled
by Bandis, 1980. In Table 16.2, just the average values
have been reproduced.
As will be noted for moderately weathered and fresh
sandstones, siltstones and limestones, the third cycle, have
Kni values in the respective ranges as follows:
Sandstone 27 to 37 MPa/mm
Siltstone 23 to 54 MPa/mm
Limestone 99 to 118 MPa/mm

Figure 16.17 Three load-unload (VJ) cycles for joints in hard


Figure 16.16 Normal stiffness testing of rock joints Vt  total unweathered rock. Bandis et al., 1983. These will be
deformation VJ  net joint closure. First cycle of used for estimating normal stiffness dependence on
load-unload. Bandis, 1980; Bandis et al., 1983. normal stress, at the highest stress levels.
Joint stiffness and compliance and the joint shearing mechanism 495

possible range of perhaps 0 to 1 MPa, and at the same 16.3.2 Normal stiffness at elevated
time correspond to the pre-consolidated condition, since normal stress levels
measured after three load-unload cycles.
From 0 to 5 MPa, there is a rapid rise of the average
(pseudo-static) normal stiffnesses to between 250 and
500 MPa/mm, and between 0 and 10 MPa, 400 to
800 MPa/mm is more common for the average normal
stiffness of the fresher joints, and about 200 MPa/mm for
the (JCS  40–50 MPa) joints in weaker rock. The
average normal stiffnesses between 0 and 50 MPa are of
the order of 2,000 to 3,000 MPa/mm for the hardest
joints.
However, considering local or incremental stiffnesses,
at tens of MPa stress levels (such as 10 to 40 MPa), nor-
mal stiffnesses for these unweathered hard joints with
high JCS may rise to tens of thousands, but may only rise
to about 1,000–2,000 MPa/mm in the case of the weaker
rocks (e.g. with JCS  40–50 MPa).
Before proceeding to a detailed treatment of possible
Figure 16.18 Comparative total deformation (Vt  rock 
joint) for different rock types, based on the most
stiffness ratios, we will present normal stiffness data
deforming first cycle of loading. Bandis, 1980. kindly prepared by Bandis from his 1980 doctoral studies
in the University of Leeds. This can be compared, where
Table 16.2 Bandis, 1980 data for Kni (3 cycles) and maximum closure Vm (3 cycles).

Vm and Kni data

1st cycle 2nd cycle 3rd cycle


Rock type and
weathering state JCS No. Kni Vm Kni Vm Kni Vm
of joints (MPa) of joints MPa/mm mm MPa/mm mm MPa/mm mm

SLATE
Fresh 175 3 35.0 .039 181.1 .027 266.2 .027
Mod. weathered 142 1 13.1 .106 69.0 .046 235.5 .039
Weathered 77 3 12.7 .331 32.6 .146 63.6 .118
DOLERITE
Fresh 167–182 2 24.2 .101 67.1 .053 111.2 .067
Weathered 60–76 3 10.8 .484 50.0 .150 80.4 .109
LIMESTONE
Fresh to slightly 152–170 11 18.8 .091 85.3 .042 117.9 .030
weathered
Mod. weathered 94–120 5 30.1 .116 51.7 .060 98.9 .040
Weathered 35–53 5 8.5 .373 44.2 .115 56.9 .079
SILTSTONE
Fresh 105 5 18.9 .135 43.9 .072 53.5 .063
Mod. weathered 67 2 10.8 .310 20.8 .205 22.9 .184
Weathered 44 3 10.4 .514 27.8 .104 35.0 .113
SANDSTONE
Fresh to slightly 68–95 8 12.8 .170 23.4 .075 37.2 .054
weathered
Mod. weathered 64–58 9 9.3 .240 17.9 .101 26.7 .089
Weathered 22 4 3.1 .469 11.3 .131 15.6 .080
496 Rock quality, seismic velocity, attenuation and anisotropy

possible with data in Table 16.4 from Pyrak-Nolte sandstone and siltstone displayed normal stiffnesses of
et al., 1990 that was reviewed in Chapter 15. between 200 and 900 MPa/mm at normal stress levels of
To conserve space in Table 16.3 normal stiffness values 10 and 25 MPa.
for weathered joints are excluded, in view of their lesser Several of the pseudo-static values of normal stiffness
relevance for reservoir conditions. A rough (JRC  15), from the tests of Bandis, 1980 shown in Table 16.3 had
weathered (JCS  44 MPa) joint in limestone showed Kn higher values of pseudo-static normal stiffness than the
values of 250, 1,275 and 2,550 MPa/mm at normal dynamic normal stiffnesses of this more deformable joint
stress levels of 10, 25 and 40 MPa. More planar in quartz monzonite. Another series of tests by Pyrak-
(JRC  6), weathered joints (JCS  25–45 MPa) in Nolte and co-workers, using a stiffer (E 32) sample, and

Table 16.3 Normal stiffnesses in (MPa/mm) measured by Bandis, 1980 in pseudo-static loading. Aperture (a) represents the approximate
unstressed aperture prior to testing. (Bandis, 2005 pers. comm.)

Limestone Limestone Sandstone Sandstone Siltstone

Normal F SW F MW F
Stress JRC  10 JRC  7.6 JRC  12 JRC  7.5 JRC  9
JCS  160 MPa JCS  160 MPa JCS  68 MPa JCS  44 MPa JCS  105 MPa
a  0.25 mm a  0.2 mm a  0.25 mm a  0.25 mm a  0.15 mm
(MPa) (MPa/mm) (MPa/mm) (MPa/mm) (MPa/mm) (MPa/mm)
10 1,200 850 470 350 350
25 6,300 10,500 12,750 12,750 3,100
40 15,750 31,500 – – –

Table 16.4 Comparison of static and dynamic normal stiffness data for the Pyrak-Nolte et al., 1987b jointed sample E 35, of 52 mm
diameter, which was the most deformable of three joint samples, possibly due to greater roughness.

Static stiffness Dynamic stiffness


Normal Stress Kn(static) Kn(dyn) Ratio

MPa MPa/mm MPa/mm Kn(dyn)/Kn(static)

2.9 1,000 4,500 4.5


10.0 2,200 8,000 3.6
33.0 3,300 25,000 7.6

Table 16.5 Selected dynamic normal and dynamic shear stiffnesses from tests on two joints in quartz monzonite. Pyrak-Nolte et al., 1990.

(all stiffnesses in (all stiffnesses


STRESS Sample E 32 MPa/mm) Sample E 35 in MPa/mm)

n (MPa) Kn(dynamic)-dry Kn(dynamic)-saturated Kn(dynamic)-dry Kn(dynamic)-saturated

2.9 15,000 35,000 4,000 9,500


10 – 80,000 11,500 20,000
20 – 100,000 20,000 25,000
70 120,000 – 32,000 59,000

n (MPa) Ks(dynamic)-dry Ks(dynamic)-saturated Ks(dynamic)-dry Ks(dynamic)-saturated

2.9 3,500 – 1,900 –


10 9,500 – 4,800 –
20 17,000 – 6,200 –
70 55,000 – 7,400 –
Joint stiffness and compliance and the joint shearing mechanism 497

the same E 35 sample as above, provided both dry and Table 16.6 Ratio of dynamic Kn/dynamic Ks for the three Pyrak-
saturated results. These are repeated in Table 16.5 in Nolte et al., 1990 joint samples in quartz monzonite.
compatible (MPa/mm) units, for ready comparison. The authors’ data set is partially incomplete. All
samples were of 52 mm diameter.

STRESS Sample E 30 Sample E 32 Sample E 35


16.4 Ratios of Kn over Ks under
Kn(dyn)/Ks(dyn) Kn(dyn)/Ks(dyn) Kn(dyn)/Ks(dyn)
static and dynamic conditions
n (MPa) (dry) (dry) (dry)

Due to equipment capacity limitations, when Bandis, 2.9 1.6 4.3 2.1
1980 tested both the normal and shear stiffnesses of the 10 1.3 – 2.4
same joint sample, involving tests in a shear box rather 20 2.0 – 3.2
33 2.7 – –
than in a high-capacity loading frame, significantly
70 – 2.2 4.3
lower levels of normal stress than any of the above were
applied. Values of Ks ranging from 1 to 7 MPa/mm were
obtained, where normal stress values were in the range
of approx. 1 to 5 MPa, for the 100 mm long samples.
It seems reasonably certain that Ks static is likely to be 16.4.1 Frequency dependence of
some orders of magnitude lower than Ks dynamic at fracture normal stiffness
comparable normal stress levels. Future testing to develop
an improved data base is obviously required. Significantly The dynamic stiffness data for the three joints reported
we do not yet seem to know the effect of in situ block by Pyrak-Nolte et al., 1990, were acquired over a lim-
sizes on Ks dynamic. Block size has a dramatic, and well ited ultrasonic frequency range (mostly 0.1 to 1 MHz).
documented reducing effect on Ks static (e.g. Barton, Different aspects of these tests were reviewed in Chapters
1972, Bandis et al., 1981, Barton, 1982), as also seen in 10 and 15, and have been summarised in part in this
Figures 16.14 and 16.15. chapter, for comparison with static data. An important
When the ratio of Kn/Ks (both static) could be com- ‘detail’ is the fact that the dynamic stiffness, showing
pared directly on the same 100 mm long samples, roughly four to eight times greater magnitude than the
mostly at the low normal stress of 1 MPa, Bandis found static values for the ‘soft’ joint in Table 16.4, is actually
that the ratio varied from extremes of 5.7 (dolerite) to dependent on the frequency.
132 (slate), but with most ratios of Kn/Ks in the narrow Pyrak-Nolte and Nolte, 1992 addressed this import-
range of 11 to 15. ant aspect, suggesting that the frequency-dependence
Table 16.6 gives the ratios of Kn(dyn)/Ks(dyn) for the may be a simple consequence of the fracture or joint-
dry samples listed in the Table 16.5. Note that the wall geometry. They referred to a ratio of dynamic/static
ratios of dynamic stiffness are for 52 mm diameter speci- normal stiffness ‘of typically three’, and likened this
mens. Under pseudo-static testing, the size of sample ratio to the difference of the dynamic to static moduli.
plays a significant role in reducing the shear stiffness, As we have seen in Part I, the ratio is rock quality
thereby increasing the ratio Kn/Ks to much larger values dependent, and the Table 16.4 result for the more
than the range given in Table 16.6. Under dynamic deformable joint (E 35), is consistent with this picture:
testing this appears not to be the case. the ‘softer’ the joints, the greater will be the difference.
It may be noted from the above tables that the range This applies to an even greater extent if the shear stiff-
of Kn(dyn)/Ks(dyn) ratios for these three, 52 mm diameter ness of the joints is also involved in the respective static
jointed samples in quartz monzonite was only from 1.3 and dynamic loading directions.
to 4.3 with a mean of 2.5. This means that the ratio of Pyrak-Nolte and Nolte 1992 applied the displacement-
the inverted compliances ZN/ZT for these small samples discontinuity theory of Schoenberg, involving assump-
ranges from 0.2 to 0.8, with a mean of 0.4. The remain- tions of discontinuous displacement, but continuous stress
ing question is what relevance this convenient ‘core- across the fractures. An assumption of inverse propor-
sized’ data has to in situ reservoirs in general, with a tionality between the discontinuity displacement and
typical spread of jointed block sizes from perhaps the specific stiffness of the fracture is also involved. The
200 mm to 5,000 mm, and mostly sedimentary rock as authors assumed that the different parts of any given
opposed to Stripa ‘granite’, or hard quartz monzonite. fracture in intimate interlock due to normal stress, will
498 Rock quality, seismic velocity, attenuation and anisotropy

earlier frequency-independent values of 11,000 and


24,000 MPa/mm (shown by black dots). For the sake of
completeness, Figure 16.19a shows one of the experi-
mental data sets for the static normal stiffness of the three
joints, which was reviewed in Chapter 10.
Pyrak-Nolte and Nolte, 1992 observed that the
dynamic stiffness is equal to the static stiffness at low
frequencies. This is clearly an important ‘end-member’
when considering the interpretation of fracture stiff-
nesses at much lower seismic exploration frequencies. It
suggests that fracture properties, as defined in detail in
Bandis 1980 static data (i.e. JCS and JRC, and result-
ant behavioural trends) will have more to say when com-
paring static and dynamic behaviour at seismic
frequencies, than when comparing static and dynamic
behaviour at laboratory ultrasonic frequencies. The in
situ scale of fracturing and block sizes, and the use of
lower frequencies, suggest that one should remain open
to the possibility that static behaviour can be a useful
guide to possible joint-parameter influences at seismic
exploration frequencies. Note from Figure 16.19 that
laboratory ultrasonic frequencies ‘automatically’ cause
Kn (dynamic)  Kn (static). At frequencies below
0.1 MHz, somewhat closer dynamic to static magni-
tudes are suggested.

16.4.2 Ratios of static Kn to


Figure 16.19 a) Static normal stiffness trends for three joints static Ks for different block
in quartz monzonite (Pyrak-Nolte et al., 1987a). sizes
b) Dynamic stiffness as a function of frequency, for
normal stress levels of 10 and 20 MPa. Pyrak-Nolte Since there are several variables involved in joint stiff-
and Nolte, 1992.
ness, beside the influence of wall strength JCS and
roughness JRC, we will in the following simplify by
assuming either small or moderate in situ block dimen-
have different characteristic frequencies, due to the dis- sions of 100 mm (i.e. bed-limited blocks) or 1 m (i.e.
tribution of voids and highly stressed rock-to-rock con- moderately jointed reservoirs). Block volumes might
tacts. This will mean that different parts of the fracture well rise to 10’s, 100’s or 1000’s of m3, in sparsely
will pass lower frequencies, while other parts will pass jointed rock masses. Since Kn is not believed to be scale-
higher frequencies. They therefore assumed that the dependent (but sample-dependent), while Ks is clearly
local transmission coefficients would depend on the very scale dependent, it can be expected that static
local static stiffnesses. ratios of Kn/Ks will tend to rise strongly with increasing
The authors used a so-called stratification percolation block size. Ks is very scale dependent because both peak
model to simulate fracture aperture distributions and to shear strength and displacement to peak are scale depend-
match (with excellent precision) experimental amplitude- ent. It could be described as doubly-dependent on scale.
frequency curves from the tests by Pyrak-Nolte et al., Comparison of the reviewed Kn trends, with the Ks
1990 referred to earlier. The result of their dynamic stiff- ranges seen at (effective) normal stresses from 1 to
ness-frequency modelling is shown in Figure 16.19b: the 10 MPa in Figure 16.14 (usually about 1 to 10
two curves apply to normal stress levels of 10 MPa and MPa/mm for 100 mm samples, and 0.2 to 2 MPa/mm
20 MPa, and were considered an improvement on the for 1 m block sizes), indicates that ratios of Kn/Ks are
Joint stiffness and compliance and the joint shearing mechanism 499

most likely to be in the following very approximate One could suggest, based on the above, that joints (or
ranges, based on data reviewed earlier: ‘meso-scale’ fractures) should be more seismically visible
than stress-aligned microcracks, or the ‘extensive dilatancy
1) 1 to 10 MPa: 100 mm size: Kn/Ks  (200–800)/ anisotropy’ (EDA), if visibility depended on compliance
(10)  20–80 contrasts, unless for some reason macro-deformation stiff-
2) 1 to 10 MPa: 1 m size: Kn/Ks  (200–800)/ ness ratios (Kn/Ks) have no influence as ‘starting points’ for
(2)  100–400 the micro-deformation dynamic compliance ratios
3) 10 to 40 MPa: 100 mm size: Kn/Ks  (1000– (ZN/ZT). In good quality rock when Estatic is close to
20000)/(10–25)  100–800 Edynamic, there should in fact be close correspondence of
4) 10 to 40 MPa: 1 m size: Kn/Ks  (1000–20000)/ Kn(static) and 1/ZN (dynamic), but apparently from what
(2–4)  500–5000 we have seen thus far, there may not be good correspon-
(Italics implies uncertainty in incremental, high stress, dence of Ks(static) and 1/ZT (dynamic), unless block size
normal stiffnesses) and frequency have influence.

This means that the pseudo-static Kn/Ks ratios may be


in the low hundreds to low thousands at typical reservoir 16.4.3 Field measurements of
effective stresses, in the case of weaker or harder rocks, compliance ZN
respectively. Here we assume that Ks remains between
about 2 and 20 MPa/mm, at large (1 m block-size), and Important recent contributions to the problem of
small-scale (0.1 m block-size) respectively, as suggested at determining field-scale joint or fracture compliances
stress levels of about 10–20 MPa in Figure 16.14. were made by Lubbe and Worthington, 2005 and
At the higher (effective) normal stress levels, Kn is seen Lubbe, 2005. These were not reviewed in earlier chap-
to rise much faster than Ks, so it is therefore that ratios of ters as they are most directly comparable with the fore-
Kn/Ks may climb to a range of about 500 to 5,000, in the going rock mechanics work concerning in situ joint
context of in situ block sizes, as opposed to joints in stiffnesses, and the scale effect known to act on the
small, highly stressed laboratory samples. It goes without (static) shear stiffness Ks. These topics belong here in
saying that the reasons for the ‘mismatch’ of these macro- Chapter 16.
deformation stiffnesses, is because they are defining such Lubbe and Worthington utilised three 40 m deep
different processes: Kn approaches the stiffness (and sta- boreholes drilled in the floor of a Carboniferous lime-
bility) of the intact rock, while Ks approaches the failure stone quarry near Bristol, in southwest England. Core
direction. (This is emphasised in Figure 15.5). analysis, wire-line logging, and local fracture mapping
These two different stiffnesses are truly the hard and provided the important characterization data, for future
the soft stiffnesses of a rock mass, just as the matrix pore comparison to other sites. Although using much lower
space, and the joints, are respectively the hard and the frequency sources than in the laboratory (2000 Hz from
soft porosities of a rock mass. However, we have seen that a sparker), these cross-hole investigations were of course
ultrasonic, laboratory-scale measurements suggest less soft- at significantly higher frequency than that of seismic
ness, since the ratios of dynamic Kn/Ks are much lower exploration. The stiffness-frequency trends shown in
than the static ratios just listed. The most important Figure 16.19 should therefore be remembered.
question that remains to be solved is whether seismic The authors found that the rock mass had a P-wave
exploration frequencies, combined with much larger in anisotropy of 10% which they attributed to the pre-
situ fracture sizes and block sizes, will cause larger dominantly horizontal, partly open fractures. P-wave
dynamic Kn/Ks ratios than those listed in Table 16.6. velocities were often in the range of 5 to 6 km/s. An
If the significantly ‘anisotropic’ or unequal static joint average of 5250 m/s for the fractured regions was quoted.
stiffnesses are in any way responsible for an (inverted) The average density was 2,600 kg/m3. For a near-surface
‘anisotropy’ or non-equality of ZN/ZT, then they are location, this velocity suggests a rock mass quality Qc
surely a strong reason for the potential ‘visibility’ of value of about 55, and a deformation modulus of about
water/oil/gas conducting joints to shear waves, as their 38 GPa, assuming a UCS of a nominal 100 MPa.
potential weakness in shear and their anisotropy of stiff- Lubbe et al., reported an intact seismic Q as high as 60,
ness (Kn/Ks) may be so marked compared to the more and an in situ value as low as 25 – the implied min-
isotropic intact rock. imum. Again the similarity of Qseis to the potential
500 Rock quality, seismic velocity, attenuation and anisotropy

static E-modulus or rock mass deformation modulus, site, i.e. crossing the basalt columns, which was probably
expressed as GPa is striking: see Figure 13.60. related to low levels of horizontal stress at this shallow
By using the upgoing or downgoing waves between test tunnel site: see Figure 13.61).
borehole source and receivers, Lubbe and Worthington Examination of Figure 16.19 and Tables 16.4 and
were able to estimate (dynamic) fracture normal com- 16.5 in fact shows the stiffer end of the above range of
pliances within the range of 2.5  1013 to in situ data to be the same as the ‘softer’ E 35 sample of
3.5  10 12 m/Pa. These we can invert to dynamic Pyrak-Nolte, when tested at only 2.9 MPa normal
normal stiffnesses in the range 4000 to 286 MPa/mm, stress, while the least stiff in situ result of 286 MPa/mm,
or reversing the order and adopting a more ‘friendly’ is obviously closer to Bandis, 1980 static Kn data shown
unit, between 0,3 to 4 MPa/
m. Assuming an average in Table 16.3.
of two continuous fractures per meter from 5 to 20 m Lubbe and Worthington considered that the vertical
depth, between all three boreholes, they estimated a fractures at the quarry (Figure 16.20) appeared to be
mean fracture compliance of 1.25  1012, which stiffer than the measured bedding planes, resulting in
converts to 800 MPa/mm in more easily understood horizontal velocities that were close to those for the
rock engineering units. Later in the paper a ‘reasonably unfractured rock. There appears to the writer the possi-
robust average’ of 1000 MPa/mm was suggested, fol- bility that the quite high horizontal P-wave velocities of
lowing some numerical scoping. typically 5.3 to 6.2 km/s could be due a horizontal stress
As they pointed out, their compliances were an order concentration beneath the floor of the quarry. If so, this
of magnitude higher (therefore the stiffnesses lower) would give an ‘equivalent depth’ correction to the empir-
than obtained from laboratory experiments at higher ical Qc – M – Vp relationships shown in Figure 13.59,
stress and on smaller and stiffer samples. However, they thereby ‘not requiring’ so high Qc values. This would
referred to other field data showing low Kn(dynamic) actually be more in line with the quite jointed appear-
values of only 500 MPa/mm (dynamic compli- ance of the quarry seen in Figure 16.20, where a rock
ance  2  10 12 m/Pa), for columnar joints in mass Q-value of about the following magnitude would
basalt, as interpreted by Myer et al., 1995 from normally be expected (see Appendix A for ratings):
Hanford Basalt cross-hole tests performed by King et al.,
1986. (In Chapter 13 we also noted a very low seismic 90 2 0.66
Q     13 (E Mass  24 Gpa)
Q in the same horizontal measurement direction at this 9 1 1

Figure 16.20 Oxford University Earth Sciences Department website photograph of the Carboniferous limestone quarry near Bristol that
was used for careful in situ investigations of normal compliance. (Lubbe et al., 2005). Note dipping bedding planes and sub-
vertical joints, presumably seldom sampled by the vertical boreholes in the floor of the quarry. The cross-hole seismic was per-
formed between three vertical holes spaced at 7 m. These are in the foreground. (www.earth.ox.ac.uk)
Joint stiffness and compliance and the joint shearing mechanism 501

The authors also discussed shear compliances at the discontinuity theory of Schoenberg, 1980, values of Kn
end of their paper, referring to Worthington and (dynamic) ranging from 22,220 to 200,000 MPa/mm
Hudson, 2000, fault studies. ‘Values of shear compli- were derived, when inverted from the reported
ance of the order of 1010 to 10 9 m/Pa can be pre- 4.5  1014 to 0.5  10 14 m/Pa compliance units.
dicted and tentatively compared to field observations’. These dynamic stiffnesses are similar, to somewhat
These can be inverted to in situ fault-related dynamic stiffer, than the Pyrak-Nolte data for natural joints of
normal stiffnesses of only 10 and 1 MPa/mm, giving a 52 mm size reviewed earlier. Lubbe, 2005 found that
somewhat closer match to the scale-dependent static the ratio of normal to tangential compliance for the
shear stiffnesses plotted in Figure 16.14 and calculated 50 mm diameter ‘fractured’ samples, was approximately
in Figure 16.15. Such in situ, modelling-based esti- 0.4 in the case of smoother surfaces, and increased from
mates would be about one order of magnitude stiffer in 0.3 to 0.8 in the case of corrugated surfaces that experi-
relation to the static stiffness values for in situ block enced ‘asperity-crushing’ beyond 30 to 40 MPa normal
sizes of say 10 m, assuming ‘in situ’ effective stress levels stress. The ratio of 0.4 for the smoother surfaces is the
above and below a nominal 10 MPa. If on the other same as the mean value reported by Pyrak-Nolte et al.,
hand, in situ block sizes were only 0.5 m, as feasible in 1990 from their tests on 52 mm joint samples, as
the neighbourhood of a fault, such ‘soft’ dynamic val- reviewed in Chapter 15.
ues (also obtained at lower seismic exploration frequen- Lubbe et al., 2005 reported details concerning the type
cies), would match static values very closely. of artificial fractures that were prepared for these tests.
Samples were either cut with a ‘fine-toothed’ diamond
saw and then polished, or, in the case of just the Jurassic
16.4.4 Investigation of normal and limestone, samples were cut and grooved to produce a
shear compliances on ‘corrugated fracture interface’ with specific voids and con-
artificial surfaces in tact areas. The two ‘fracture’ types, which were all tested
limestones dry, are shown in Figures 16.21 and 16.22, together with
the respective, and contrasting, compliance-normal stress
Lubbe and Worthington, 2005 also referred to Lubbe, behaviour. In these figures we have added stiffness scales
2005 laboratory data for artificial fractures machined on the right-hand axes, to aid comparison with the labora-
from Jurassic and Carboniferous limestones, tested dry tory stiffness data reviewed earlier in this chapter.
under 5 to 60 MPa normal stress, at ultrasonic frequen- The Jurassic limestone samples had porosities of
cies between 0.6 and 1.0 MHz. Using the displacement approximately 13%, while the Carboniferous limestone

Figure 16.21 a) Samples of Jurassic (J) and Carboniferous (C) limestone, with ‘fractures’ prepared with a fine-toothed diamond saw. b) Shear and
normal compliance versus normal stress, with inversion to dynamic stiffnesses shown on the right-hand axis. Lubbe et al., 2006.
502 Rock quality, seismic velocity, attenuation and anisotropy

samples had porosities of only 2.4%. Densities were It seems, also from earlier review of the Perspex plates
2710 and 2660 kg/m3 respectively. model of Hsu and Schoenberg, 1993, and of a cut ‘frac-
The lower stiffnesses seen in the case of the corru- ture’ in aluminium, that the less ‘rock-joint-like’ the
gated fracture surface, giving artificial (initial) limits to sample, the more likely it is that the ratio of ZN/ZT will
contact areas, has some resemblance to the reduced approach 1.0. In the case of the above grooved lime-
(static) normal stiffnesses recorded by Bandis, 1980 for stone interfaces, the ratio rose markedly, towards 0.8,
slightly sheared joints, shown in Figure 16.16 (see ‘mis- beyond a stress level of 40 MPa.
match’ result). A marked acceleration in the axial short- In the rock mechanics world of pseudo-static testing,
ening of the sample was noted by the authors between including laboratory and field scales of blocks, the more
stress levels of 30 and 40 MPa, which also corresponded ‘rock-mass-like’ the sample becomes (i.e. when moving
with the increment of dynamic normal stiffness seen in into the in situ environment), the higher the ratio of
Figure 16.22. It is also consistent that a decrement of Kn/Ks. If this behaviour has a certain influence, even only
dynamic shear stiffness should also occur at this same minor, on exploration seismic behaviour, (at orders of
stress level, resembling a ‘roughness-destroyed’ portion magnitude lower frequency than all the laboratory
of pseudo-static shearing simulation (the JRCmobilized compliance testing reviewed in this book), then a lower
concept), shown in Figure 16.10. ratio of ZN/ZT would be expected, far from the neigh-
The authors discussed the marked difference of com- bourhood of 1.0, as seen in many laboratory-based or
pliance behaviour exhibited by their two types of artifi- modelling-based geophysics publications.
cial fracture. The ‘grooved’ sample showed a steady Concerning the actual effect of water saturation com-
reduction in both normal and shear stiffness up to 30 or pared to the dry state, which we could tentatively
40 MPa, exactly the reverse of the stiffening behaviour extrapolate to the hydrocarbon equivalent of ‘brine-
of the sample that actually had a more ‘natural’ contact saturated compared to gas’, we will see later in this
area development. They proposed, very reasonably, that chapter that both the pseudo-static Kn and Ks magni-
a gradual weakening (of the asperities) was occurring in tudes will reduce somewhat when water saturated.
the grooved case. They also reported a layer of pow- Devoid of any attenuation in the form of fluid-changed
dered rock development beyond 40 MPa. Finally they scattering or squirt losses, these reductions can simply
doubted that the theoretical use of the ZN/ZT compli- be explained (and quantified) by reduced JCS (joint
ance ratio was justified, as an indicator of potential fluid wall strength) with moisture for the case of a reduced
saturation (and fluid type), as reviewed and also doubted Kn, and reduced JCS and r (residual friction angle),
in Chapter 15 (see Figure 15.11). for the case of a reduced Ks.

Figure 16.22 a) Samples of Jurassic (J) limestone, with ‘fractures’ prepared by cutting and grooving, to produce a ‘corrugated fracture inter-
face’. b) Shear and normal compliance versus normal stress, with inversion to dynamic stiffnesses shown on the right-hand
axis. Lubbe et al., 2006.
Joint stiffness and compliance and the joint shearing mechanism 503

Although the micro-displacement phenomena mobil- numbered 1, 3 and 5 extending from ultrasonic to sonic
ized by dynamic waves may not be fully affected by such to cross-hole frequencies, represented Lubbe’s D.Phil.
pseudo-static reductions in stiffness, there is of course contribution to this important compliance-size trend.
increasing evidence and logic, in reduced dynamic stiff- We will summarise the differently numbered data sets,
nesses in the case of weakened joint or fracture surfaces. and also convert compliance units to dynamic stiffness
After all, such reductions will also tend to reduce Estatic units, for easier reference to both pseudo-static and
(or M), and Edynamic, assuming that frequencies are not dynamic test data, where the ‘Pa/m’ format (or
too high to prevent micro-flows of fluid. MPa/mm) is used, instead of the geophysicists’ some-
what cumbersome ‘m/Pa’, with the need of multiplica-
tion by a very small number like 1012.
16.4.5 The Worthington-Lubbe-
Hudson range of compliances 1. Laboratory compliance tests, using artificial 50 mm
diameter surfaces in Jurassic and Carboniferous
Lubbe, 2005 considered that the main conclusion from limestone. The vertical bar bounds the 5 to 60 MPa
his research on compliance, could be summarised in normal stress range. The complete range of both
Figure 16.23. This was kindly provided for reproduc- normal and shear compliances was 0.5  1014 to
tion in this book. The careful three-author title to this 4.5  10 14 m/Pa, or 200,000 to 22,220 MPa/mm
sub-section is due to the joint contribution of these (equivalent to a very stiff-sounding 200 to
researchers, and is also based on the figure’s earlier 22.2 MPa/
m). See Figures 16.21 for the differen-
appearance in Worthington and Lubbe, 2004. The data tiation of ZN and ZT results. (Data from the
‘grooved’ samples: Figure 16.22, are not incorpo-
rated in Figure 16.23).
Kn (dyn) (MPa/mm)
Normal or tangential fracture compliance (m/Pa)

10-9 6
1 2. The compliance interpretations of Pyrak-Nolte et al.,
1990 for dynamic tests on three natural joints in
10 quartz monzonite, of 52 mm diameter. Normal
10-10
stress levels ranged from 1.4 to 85 MPa. Compli-
100 ances ranged from 5.3  1013 (the lowest ZT
10 -11
4
5
value) to 10 14 m/Pa (the highest ZN value). In
2
3 1000 stiffness units, these correspond to 1,900 MPa/mm
1.4 MPa
to 100,000 MPa/mm (or 1.9 to 100 MPa/
m).
10-12 3. This sonic-log based field data was from the
10,000
1
5 MPa Carboniferous limestone quarry at Tytherington,
60 MPa 85 MPa
100,000
depicted in Figure 16.20. The approx. 23 kHz
10-14
10-2 10-1 100 101 102 103 compressional wireline measurements were analysed
CORE LOG XHOLE
Log fracture dimension scale (m)
SEISMIC
by Lubbe, first using a ZN  ZT assumption and
the theories of Schoenberg, 1980 and Pyrak-Nolte
Figure 16.23 Scaling of fracture compliance from laboratory and et al., 1987b. Allowance for a 20° variation in frac-
field data. The data numbered 1, 3 and 5 was pro- ture dips resulted in an estimate of ZN  4.8
duced by Lubbe from compliance measurements on (2.6)  1013 m/Pa, or 2,080 MPa/mm (range
artificial fracture surfaces in limestone, reviewed 4,405 to 1,350 MPa/mm). Lubbe also used the
above, from the limestone quarry data reviewed ear- Schoenberg and Sayers, 1995 excess compliance the-
lier, and from a highly fractured shale-rich field site. ory, using the vertical velocity reduction accorded
The latter two included sonic logging and cross-hole
to fracturing (500 m/s) to give a 2.5  1013
interpretations, assuming a given mean bedding-
 ZN  5.0  1013 m/Pa estimate of normal com-
plane fracture spacing. The other data are explained
in the numbered text that follows. The horizontal
pliance. This range converts to a Kn (dynamic)
axis of the figure was diagrammatic, since ‘it is not range of 4,000 to 2,000 MPa/m respectively. Lubbe
possible to measure any parameter that accurately used a third calculation method, dispensing with
represents the size of the fracture’. (Worthington the ZN  ZT assumption, and basing his estimates
and Lubbe, 2004, Lubbe, 2005, Figure 7.1 and on Hudson and Crampin theories to interpret the
Lubbe, 2005 pers. comm.). 10% P-wave anisotropy as an assumed 0.035
504 Rock quality, seismic velocity, attenuation and anisotropy

fracture density for the horizontal ‘open’ joints. This 10.67). The shear compliance estimate was
resulted in a similar 2.6  1013  ZN  5.2 1.1  109 m/Pa, or a dynamic shear stiffness of
 1013 m/Pa, or 3,850 to 1,920 MPa/mm 0.9 MPa/mm. The theories of Hudson and
dynamic normal stiffness. As one may note from Schoenberg were used to derive a range of possible
the location of these larger scale, lower frequency low stiffnesses, as reviewed in Chapters 10 and 15.
data in Figure 16.23, only the low stress laboratory The frequency band was 10 to 150 Hz.
ultrasonic data lies in the same magnitude bracket. 7. Finally, the dotted black line in Figure 16.23 was
4. The data point labelled ‘4’ in Figure 16.23 was the based on the theoretical prediction by Hudson et al.,
Myer et al., 1995 normal compliance estimate for 1997 for tangential compliance. For fault-like fea-
Hanford Basalt columnar jointing, derived from tures it also applied to normal compliance if the
cross-hole measurements. The normal compliance ‘fracture’ was filled with weak material, assuming a
estimate was 2  1012 m/Pa, or a dynamic normal 10% area in ‘welded contact’. Lubbe, 2005 therefore
stiffness of 500 MPa/mm. (In Worthington et al., considered that the internal consistency of this spread
2001, the respective values are 5  10 12 m/Pa, or of data demonstrated that compliance increased with
200 MPa/mm). Both are of similar magnitude to the assumed scale of the fractures, dynamic stiffness
the laboratory (100 mm) scale pseudo-static Kn thereby reducing with size. The experience from
data obtained at higher (10 MPa) stress, as meas- pseudo-static testing, and also from structural-geo-
ured by Bandis, 1980 (see Table 16.3). logical logic, obviously supports such a thesis, and
5. The vertical bar labelled ‘5’ in Figure 16.23 is from has long been a part of shear stiffness prediction in
a highly fractured shale-rich field site called relation to block size, since Barton, 1972.
Reskajeage, where Lubbe, 2005 again used Hudson
and Crampin theories to relate a 22% P-wave Lubbe, 2005, perhaps incorrectly, felt that the very
anisotropy to a 0.07 fracture density. Normal com- low interpreted shear compliance for the North Sea
pliance estimated for single fractures were as fault zone might be highlighting ‘the theoretical possi-
follows: 6.6  1013  ZN  8.3  1013 m/Pa, bility of ZN/ZT being close to zero’, since he considered
which converts to a dynamic normal stiffness range that the normal compliance of a fracture at 2 km depth
of 1,510 to 1,208 MPa/mm. A second method of ‘would be vanishingly small’. Clearly this is an exagger-
estimation, again involving the ZN  ZT assump- ation, as the effective stress at 2 km depth in a petrol-
tion and Schoenberg and Pyrak-Nolte theories, eum reservoir may be ‘only’ 20–25 MPa. However, the
resulted in the following range of compliances: important point was emphasised that the ratio ZN/ZT
1.9  1012  ZN  2.4  10 12 m/Pa, meaning could be very small, thereby indirectly linking the dynamic
dynamic normal stiffnesses ranging from 525 to behaviour to the pseudo-static behaviour. In sub-
410 MPa/m. Thirdly, Lubbe, 2005 followed the section 16.4.2, potential pseudo-static ratios of Kn/Ks
methodology of Pyrak-Nolte et al., 1990 and Myer for typical 1 m in situ block sizes were derived, ranging
et al., 1995, modelling the observed attenuation from 100–5000. These are large ratios, but they are finite.
from the P-wave transmission amplitudes in a ZN/ZT never vanishes, as perhaps implied in Figure
highly anisotropic ‘open’ fractured zone, deriving a 15.10, at very small aspect ratios. That ZN/ZT becomes
normal compliance of 4.5  1013 m/Pa, or a very small at large scale is however easy to understand.
dynamic normal stiffness of 2,230 MPa/mm, both Worthington and Lubbe, 2004 presented a slightly dif-
reportedly consistent with cross-hole travel time ferent version of Figure 16.23, describing their analysis of
data. Lubbe remarked that the Reskajeage compli- a cross-hole tomography investigation in Cornwall,
ances were higher (and stiffnesses therefore lower) which was published by Herwanger et al., 2004. They
than at the limestone quarry site, due to the greater estimated the compliance contribution of a dominant,
continuity of hydraulically conducting ‘open’ dipping fractures zone, which showed velocities with a
fractures. range of about 3.0 to 4.5 km/s, and an estimated frac-
6. The data point labelled ‘6’ in Figure 16.23 was ture density of 0.07. Using the Schoenberg and Sayers,
from Worthington and Hudson, 2000 and related 1995 excess compliance concept, they estimated total
to a computed shear compliance estimated for a values of ZN and ZT for the fracture zone of
major fault zone from 1 to 2 km depth, intersecting 6.5  1012 and 7.0  1012 m/Pa. Due to a hetero-
a North Sea well at an angle of 50°. (see Figure geneous distribution of fracture frequency in the zone,
Joint stiffness and compliance and the joint shearing mechanism 505

Table 16.7 An assembly of (macro-deformation, pseudo-static)


normal and shear stiffnesses from laboratory and in
situ tests on clay-filled discontinuities, from Brazilian
dam foundations. After Infanti and Kanji, 1978.

Clay-filling Ks Kn Ratio
Thickness (mm) MPa/mm MPa/mm Kn/Ks

50–100 0.01–0.1 0.1–0.5 6 (5 to 10)


12–20 0.1–0.6 0.5 – 2.0 3.6 (3.3 to 5)
1 mm 1.0 5.0 5

they estimated a possible range of 7.0  1013 m/Pa if


10 fractures/m, and 7.0  1012 m/Pa if only 1
fracture/m. These invert to dynamic stiffnesses of 1,430
and 143 MPa/m respectively. Their estimated range
could be added as a ‘5b’ set of data in Figure 16.23
(immediately to the right-side of line ‘5’, thereby
slightly extending the trends for lower stiffness at larger
assumed size, as sampled by ‘lower’ frequency.

16.4.6 Pseudo-static stiffness data


for clay filled discontinuities
and major shear zones

While addressing the subject of likely large-scale com- Figure 16.24 General trends from in situ tests at dam sites, concern-
pliances and dynamic stiffnesses, civil engineering con- ing the (pseudo-static) shear stiffness of normal-and-
tributions to the subject of pseudo-static stiffnesses of shear loaded rock joints and clay-filled discontinuities,
clay-filled (i.e. major) discontinuities can be cited, from as a function of effective normal stresses in the near-
Brazilian dam foundation testing. Table 16.7 shows surface range of 0.1 to about 4 MPa. I) Rough unfilled
macro-deformation normal and shear loading tests joints, or with thin clay film, peak  1 mm. II) Joints
from the 1970s, reported by Infanti and Kanji, 1978. without filling, or with thin clay film, peak
1.0–2.5 mm. III) Filled joints (thickness up to
The applied effective normal stress levels were from
12 mm) with peak 2.5–5.0 mm. IV) Smooth joints
only 0.5 to 2.5 MPa, representing ‘near-surface’ condi-
with fillings 20 mm, and peak  20 mm. Infanti
tions in a geophysics perspective. and Kanji, 1990.
A much more comprehensive assembly of shear stiff-
ness data from in situ testing at Brazilian, and other
dam sites was subsequently reported by the same injection or pumping could be monitored by so-called
authors, in Infanti and Kanji, 1990. The general trends ‘Pac-ex’ combined double-packer-extensometer units
for their extensive shear stiffness data are summarised in installed in boreholes intersecting the fracture zones.
Figure 16.24. Note the dominance of values of shear Interpretation of the normal-deformation effective-stress
stiffness of less than 1 MPa/mm, but at stress levels that change allowed estimation of large scale pseudo-static
are an order of magnitude less than reservoir effective normal stiffness from measurements in a total of ten
stresses. Much of the data range corresponds to the boreholes, with numerous measurements concentrated
black-discs shown at 0.7 m size in Figure 16.14. in ‘Fracture zone 2’, which was a major dipping feature
An interesting coupled-process investigation of large intersected over hundreds of meters extent at the URL.
scale fracture zones at the Underground Research The data sets describing the in situ normal stiffnesses
Laboratory (URL) in Manitoba, based on extensive were extensive, due to the range of normal stress levels
borehole-based measurements, was reported by Martin measured in different locations adjacent to and within the
et al., 1990. Hydraulic pressure changes caused by fracture zone, and due to the fact that the pore pressure
506 Rock quality, seismic velocity, attenuation and anisotropy

could be altered at will. In a so-called cataclastic and on a planar fracture actually caused normal incidence
very permeable section of FZ 2, where the effective P:S conversion. This led Schoenberg and Nakagawa,
stress was determined to be as low as 0.2 to 2 MPa, the 2002 to demonstrate that linear slip theory and off-
measured normal stiffnesses ranged from only 2 to diagonal terms in the Schoenberg fracture compliance
6 MPa/mm. Elsewhere, where effective normal stresses matrix, could be used to relate the dynamic traction on
were as high as 20 MPa, the normal stiffness ranged the experimental fracture with the dynamic displace-
between about 20 and 500 MPa/mm. Flow-rate inter- ment discontinuity across the fracture.
preted hydraulic apertures were as large as 400
m in the They found that due to an applied shear stress, a frac-
cataclastic zone, and as small as approx 20, 40 and ture compliance matrix had a single off-diagonal term
60
m where normal stiffnesses and stresses were an that coupled the normal component of dynamic stress
order of magnitude, or more, higher. Reasonably good with the tangential component of displacement, and
correspondence to Barton-Bandis coupled modelling vice versa. The off-diagonal component (ZN V) appeared
was reported by the authors. from their laboratory tests to be linearly related to the
The low values of normal and shear stiffness reviewed shear stress component.
here, represent very attenuating, heavily fractured, and In a partly related field, where joint stiffnesses are of
sometimes clay-bearing rock masses. In terms of attenu- concern for the modelling of deformation in rock
ation (1/Q) they would represent cases of complete masses as a result of, for example, tunnel or slope excav-
energy loss, not per cycle, but in just part of a cycle, i.e. ation, we generally assume a linear initial shear stress-
below the traditionally defined Qseis  2 ‘limit’. displacement relation. If the joints are displaced in shear
As shown earlier in this chapter, ratios of pseudo-static (as depicted in Figure 16.16 – see ‘mismatch’), there will
Kn/Ks perhaps of several hundred to several thousand, are tend to be a linear relation between normal stiffness and
suggested if clay-free fracturing or jointing is to be normal stress, following Bandis, 1980 and Bandis et al.,
represented. But with a dominance of clay-filling due to 1983. Perhaps from these physical results we can
faulting as above, ratios of Kn/Ks perhaps as low as about approximate the off-diagonal response referred to above.
5, and therefore ratios of ZN/ZT as high as about 0.2, An example of the Barton-Bandis shear strength-
would be suggested. (Refer also to the discussion on fault- displacement-dilation modelling of joints or fractures of
scale stiffnesses given in Chapter 15, section 15.6.1). different sizes, is shown in Figure 16.25. One may note
Of course it is not known whether the existing ‘static’ the modelling of a delayed dilation as block size increases.
stress-deformation conditions acting on the joints or frac- Each of the responses to shear stress are scale-defined. The
tures in question, could be considered to have signifi- likelihood of minor cross-jointing in situ, to thereby
cant influence on the stiffer dynamic micro-displacement define blocks of a certain size, will be found to give differ-
‘excursions’. Such cyclic excursions presumably will ent macro-shear stiffnesses, due to the ‘double-effect’ of
tend to occur with suitably increased gradient, above scale on both peak shear strength and displacement to
the existing in situ stress-deformation gradient (or load- peak; the latter also tending to determine the initial slope.
ing path). When this is already ‘soft’ due to clay-filling, The extent to which this ‘macro-deformation’ behav-
higher (dynamic) compliances seem reasonable to iour will ‘steer’ dynamic micro-deformation behaviour
expect. Interestingly, Lubbe, 2005 refers to Japanese is of course uncertain, but the scale effect seems unlikely
work concerning faults under high stress that were to be totally ignored. As a corollary, why should small
assumed to potentially suffer inelastic behaviour in the laboratory samples give more ‘correct’ values of compli-
solid wall-rock material, during transmission of elastic ance and stiffness than in situ block sizes? Possibly the
waves within the shallow sub-surface. ‘Reverse’ stiffen- Worthington-Lubbe-Hudson data shown in Figure
ing, as in the grooved sample of Lubbe (Figure 16.22) 16.23 is tending to resolve this dilemma.
could then affect the compliance/stiffness response. In view of the Nakagawa laboratory observation that
application of shear stress causes P:S conversion on a
fracture, one may speculate whether the ‘open’ fractures
16.4.7 Shear stress application may that are the dream of well production, are also under
apparently affect compliance sufficient shear stress through limited mis-alignment
with major principal stress, that the shear wave splitting
Nakagawa performed an interesting laboratory test which phenomenon is also affected in an equivalent manner
demonstrated that the application of static shear stress to the above wave conversion.
Joint stiffness and compliance and the joint shearing mechanism 507

excursions in this area, it is surely justified to continue


the classic assumption that ‘friction’ is indeed one of the
sources of seismic attenuation (Q1) in a rock mass.

16.5 Effect of dry or saturated


conditions on shear and normal
stiffnesses

The effect of the environmental conditions (gas, brine,


or oil) on the anisotropy-causing compliances of rock
joints or fractures, at levels of effective normal stress rele-
vant to reservoir conditions, have obviously not yet been
investigated to the level required, now that shear wave
splitting and fracture diagnostics are becoming more com-
mon, and more important in hydrocarbon exploration
and time-lapse monitoring. There is the added complica-
tion that the application of high, reservoir-related stress
levels has demanded the use of rather small (roughly 40
to 50 mm) sample sizes in the very few laboratories who
have investigated (dynamic) joint compliances. We have
reviewed readily available data, under various contexts,
in Chapters 10, 13 and just now in this chapter.
With the exception of temperature and oil-saturation
effects, there is a large body of relevant test data from
earlier rock mechanics investigation that can be used to
deduce the approximate normal and shear stiffness
magnitudes for rock joints. Joints and fractures developed
in different rocks obviously have different rock
strength, wall-alteration and roughness (UCS, JCS and
JRC) characteristics. Although not direct models for
compliance, the different characteristics can perhaps be
Figure 16.25 Example of manually calculated shear stress- of value in guiding the extrapolation of the presently
displacement-dilation curves for different block sizes, very small, and very small-scale compliance data set.
using the JRCmobilized concept of Barton, 1982. Scaling The variable shear strength and shear and normal
of parameters JRC and JCS shown in the table, fol- stiffnesses of rock joints have been of concern for a long
lows developments described by Bandis et al., 1981. time (for almost 40 years, at publication time), due to
the need for basic rock stability calculations, and for
subsequent coupled flow-deformation analyses.
Figure 16.10 showed the dimensionless JRC mobil- It is therefore logical to look at this earlier evidence
ized concept that is used to predict the various ‘macro- for the effect of dry (assume equal to gas-saturated) or
deformation’ shear strength-displacement-dilation water-saturated conditions on joint or fracture proper-
behaviours depicted in Figure 16.25. In the initial part ties. This line of investigation is followed here, under
of the curve where ‘mobilization of friction’ is written full recognition that the ‘macro-deformation’ behaviour
(i.e. a reservoir joint under limited shear, or differential of rock joints has to be applied with caution to the ten-
stress), one may imagine an impinging dynamic wave tative extrapolation of micro-deformation dynamic/
causing very slight, cyclic,
m-magnitude (?) shear dis- seismic wave loading effects, being mindful of the
placements. Since such shear waves are known to usual Edynamic  Estatic inequality.
‘mobilize’ a shear compliance of recognisable magni- Before referring back to ‘environmental’ test data
tude with recognisable units (m.Pa1) during dynamic from the 1960s and 1970s, it is helpful to address the
508 Rock quality, seismic velocity, attenuation and anisotropy

components used in the most comprehensive-yet-simple-


to-apply joint model. The Barton-Bandis criterion is
widely used and referenced outside geophysics litera-
ture, where it is unknown, and is part of a well estab-
lished rock mechanics numerical modelling procedure
(UDEC-BB). A useful feature of the model is the ease
with which input data can be acquired from index tests
on drill-core.
The shear-related components JRC, JCS and b
were introduced by Barton, 1973a, and were refined by
the addition of r and simple index test methods by
Barton and Choubey, 1977. This development was
based on numerous tilt-shear and direct shear tests on
130 fresh and partly weathered joint samples. The just
developed JRC-JCS-r joint strength parameters were
subsequently utilised by Bandis, 1980 in very extensive
studies of both shear and normal stiffness and scale
effects on shear strength. These three strength param-
eters, and the scaling of the first two, have been referred
to elsewhere in this chapter. Here we will provide some
details of the parameters themselves.

16.5.1 Joint roughness coefficient


(JRC)

It is reasonable to believe that the joint roughness coef-


ficient (JRC), which varies from 0 to about 20 (for
Figure 16.26 Laboratory-scale joint roughness profiles with their
completely plane to extremely rough joints) is not sig-
measured JRC values, from Barton and Choubey,
nificantly affected by the dry or wet condition, since it 1977. Note that the roughness is defined by one
is essentially ‘geometry’. Examples of laboratory-scale parameter for simplicity. The lab-scale JRCo value is
(100 mm) profiles and their JRC values (now termed reduced when larger block sizes are involved, using a
JRCo) are shown in Figure 16.26. Examples of JRCo to JRCn scaling equation, for block sizes of
JRCn profiles of longer joints, as may be used in length Ln in place of Lo. Scaling equations are given
larger-scale shear strength estimation, are shown in later in the text.
Figure 16.27.
In softer, younger rocks, JRC may perhaps reduce
slightly during the process of compaction caused by without the writers knowledge, has called ‘the Barton
pore pressure reduction when producing from a reser- comb’. The amplitude (a) of asperities per measure-
voir. This effect, which can hardly be evaluated, is a ment length (L) can also be used, with a/L scaling to
function of the changing effective stress/strength ratio: JRCn (Barton and Bandis, 1990).
specifically /JCS.
n In practice, the index tests illustrated in Figure 16.28
The idea behind JRC is that it is extremely easily are performed on several samples of each joint set that
determined, as indicated by several index test methods, is to be modelled (e.g. in a UDEC-BB distinct element
shown diagrammatically in Figure 16.28. These include model). Makurat et al., 1990 gave examples of such
self-weight tilt tests which can be performed on joints applications, from one of the earlier tunnel modelling
recovered from core, as for instance performed on exercises with UDEC. An extract from this modelling,
joints from the Ekofisk chalk reservoir in the mid- concerning the different apertures (e and E), their dis-
eighties. Also shown is the roughness-profile gauge tribution with depth, and effects of tunnel excavation,
‘comb’ method, which one enterprising manufacturer, was illustrated in Figure 15.50.
Joint stiffness and compliance and the joint shearing mechanism 509

16.5.3 Basic friction angle Fb and


residual friction angle Fr

The third and fourth components of the shear strength


criterion are the basic friction angle (b) of unweathered
artificial, planar, dry rock surfaces, and the residual fric-
tion angle (r) applying to flat, non-dilatant, saturated,
well-sheared surfaces that may be weathered, i.e.
r  b. Since it is very difficult to reach this minimum
strength value in standard direct shear tests, an ultra-
simple empirical equation was developed to enable esti-
mation of r from the easily measured b (often close to
30°), using Schmidt hammer rebound (r) on the satu-
rated joint walls to compare to the rebound (R) on the
dry artificially cut rock surfaces. With r  R, the residual
strength is estimated as follows:

r  ( b 20) )  20(r/R ) (16.3)

Figure 16.27 Large-scale (1 m) laboratory-determined roughness The back-calculation of JRC from tilt tests or direct
profiles, and their JRCn values as determined from shear test, and the JRC-labelled profiles in Figure 16.26
self-weight tilt testing, from Bakhtar and Barton, 1984. are based on this method of estimating r . Since mois-
ture generally reduces both JCS and r, there are
inevitable consequences of saturation both on the shear
stiffness (which is given by JRC, JCS and r), and on
16.5.2 Joint wall compression the normal stiffness (which is given by JRC and JCS).
strength (JCS) Barton, 1973a and Barton and Choubey, 1977, gave
extensive reviews of the effects of saturation on b, r,
The joint wall compression strength (JCS) is known to and c. One of the first sources of data for these reviews
generally reduce with water saturation compared to the was Horn and Deere, 1962, who showed that massive
dry state. This is because of the much researched effect crystal structures such as quartz and calcite may cause
of moisture on the uniaxial compressive strength ( c), the ‘coefficient of friction’ (flat, artificial surfaces, i.e. b
and on the Point load tensile strength (I50), which in the context of these tests) to increase with saturation,
emanates from the late 1960s. When correctly per- while layer-lattice structures such as mica and chlorite
formed, JCS measurement by Schmidt hammer is per- caused the ‘coefficient of friction’ (b) to reduce.
formed on clamped, water saturated joint samples, Barton, 1973a, tabulated numerous test results from
specifically to record eventual moisture-reduced wall the dry and saturated states for the three categories 1) b
strength, with JCS  c in general, as illustrated in (or r) 2) c (or t: the tensile strength), and 3) the over-
Figure 16.5a. all shear strength (which incorporates effects on
It will be noted from the peak shear strength criterion b, r and JCS). Rock types included sandstones, silt-
of Barton-Choubey (for laboratory scale estimation), stones, limestones, chalk, and shale, and several igneous
and from the Barton-Bandis (larger scale) version, to be rocks such as granites, basalts, dolerites, porphyries, meta-
given as equation 16.5, that when JRCo or JRCn are morphic gneisses and slates. This review showed that:
zero, due to complete planarity and smoothness (but
without complications of work-hardened polishing), the 1. b may typically range from 26° to 34°, while r
ratio JCS/ n no longer influences the shear strength. can be several degrees less, depending on the degree
This assumes of course that the remaining value of shear of weathering (since r  R). In general terms,
strength ( n tan r) correctly captures the essence of r moisture (‘wet’ compared to ‘dry’) caused from 0°
for the rock concerned. This is the function of the fol- to 4° (generally 1° to 3°) reduction in b and r,
lowing ratio r/R. since pure crystal structures (massive/layer-lattice)
510 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.28 Diagrammatic illustration of joint characterization methods, from Barton, 1999a. a) The left column illustrates direct shear box
testing, where shear force T is designed to act ‘in-line’ in practice, to avoid over-turning moments. Note shear strength-dis-
placement-dilation and shear strength envelope results for three hypothetical joint samples. b) The second column illustrates
self-weight tilt tests, performed on the joints for back-analysis of JRC, and performed on (un-polished) core sticks or sand-
blasted flat-sawn surfaces (both unweathered) for input of b to the empirical equation for estimating r. c) The third column
shows Schmidt hammer tests for estimating JCS and c, and for input of the degree-of-weathering ratio r5/R5 (see worked exam-
ple showing use of highest 50% in each case) for input to the r estimation. d) The fourth and final column illustrates ampli-
tude/length (a/L) measurement, and ‘comb’ roughness profiling, with respective application in the a/L diagram for JRCn
estimation, or for JRCo profile-matching at laboratory scale, using Figure 16.26. e) At the bottom of three of the columns, the
statistical distribution of each of the key joint strength-and-stiffness parameters is illustrated, together with a reminder of the
likely scale-effects on JRC, JCS and the uniaxial compressive strength c. The core photograph is a fine example of contrasting
JRC values from an inter-locked joint and a minor fault in welded tuff, with respective JRCo values of about 15 and 1.
Joint stiffness and compliance and the joint shearing mechanism 511

Table 16.8 Examples of the effect of water saturation compared to Barton, 1982 for derivation of the following empirical
dry conditions, on the basic friction angles of flat, equations, two of which were derived in Bandis et al.,
unweathered rock surfaces. From review by Barton, 1981 from Bandis, 1980 scale-effect experiments).
1973a.

Rock type (flat, smooth, peak


non-polished surface) b dry (degrees) b wet (degrees)
A ) Shear stiffness Ks  (16.4)
peak
sandstone 26–35 25–33
siltstone 31–33 27–31 B ) Shear strength
limestone 31–37 27–35
  JCSn  
peak  ′n tan  JRCn log    r  (16.5)

 ′ 
 n 
are seldom the only component of rock joint sam-
ples as opposed to mineral samples. However, an
exceptional 9° reduction for slate (dominant layer- C ) Displacement to peak strength
1
lattice) was noted due to water saturation. Three L n  JRCn 
3

examples are given in the Table 16.8 extract from peak ≈   (16.6)
500  L n 
Barton, 1973a.
2. In general terms moisture (‘wet’ compared to ‘dry’) (peak, Ln in metres)
caused from 10% to 50% (generally 20 to 40%)
reduction in uniaxial strength ( c or UCS), and a D ) Large -scale joint roughness
similar reduction in point load or tensile strength
 L 0.02 JRCo
( t). (Water injection into a reservoir like chalk can JRCn ≈ JRCo  n  (16.7)
therefore have serious consequences – which may  L o 
be overridingly positive especially if planned for).
Water saturation causes reduced UCS, reduced JCS,
E) Large -scale joint wall strength
reduced b and reduced r, therefore reduced joint
normal stiffness, reduced joint shear stiffness and  L 0.03 JRCo
JCSn ≈ JCSo  n  (16.8)
reduced peak and residual shear strength.  L o 

As shown in Figure 16.28, the value of JCS can be


estimated from Schmidt hammer tests on saturated, F ) Initial unstressed aperture
JRCo  
 0.2 c  0.1
fresh or weathered joint surfaces, using a density-uniaxial Eo ≈ (16.9)
compression strength conversion, from Miller, 1965. See 5  JCSo 
also ISRM, 1978. It is logical to expect that JCS will also
be adversely affected by moisture, as is the case for UCS, where Lo  lab-scale sample length (usually
since JCS specifically involves asperity failure/ stiffness 100 mm)
when contributing to either the shear or normal stiffness Ln  in situ block size (generated by the
in the Barton-Bandis models for these two components. intersecting joint set)
Most likely JCS will be more sensitive to weakening by Eo  millimetres
water than the rock matrix, if there is any trace of weath-
ering, and because of the increased surface area and We will assume the following input data:
microcracking associated with jointing.
r dry  30° (e.g. gas) r wet  27° (e.g. brine)
JCSo dry  70 MPa JCSo wet  50 MPa
16.5.4 Empirical equations for the
JRCo  8 Lo  100 mm Ln  500 mm
shear behaviour of rock
joints n  10 MPa, 20 MPa

We can illustrate the above effects of the dry or saturated Thus according to the expected effects of moisture
state on shear stiffness, by evaluating examples. (See on two of the three strength components, there is a
512 Rock quality, seismic velocity, attenuation and anisotropy

Table 16.9 Barton-Bandis modelling of the possible effects of water saturation on the shear stiffness of joints sampled at two different scales.

A. Dry cases B. Wet cases


0.028
 500 
1. JRC n ≈ 8    6.2
 100  1. JRC n (assume unchanged)

0.038 0.038
 500   500 
2. JCSn ≈ 70    47.6 MPa 2. JCSn ≈ 50    34.0 MPa
 100   100 
1
 6.2  3
3.  peak ≈ 0.001   ≈ 2 . 3 mm (0. 0 023 m) 3.  peak (assume unchanged)
 0.5 
  47.6     34.0  
4. i) peak  10 tan  6.2 log    30)   6.8 MPa 4. i) peak  10 tan 6.2 log 
  10     10   27 )  5.8 MPa

  47.6     34.0  
ii) peak  20 tan  6.2 log  ii) peak  20 tan 6.2 log 
  20 
  30)   12.7 MPa
   20   27 )  10.8 MPa

6.8 5.8
5. i) K s ≈ ≈ 3.0 MPa/mm at n  10 MPa 5. i) K s ≈ ≈ 2.5 MPa/mm at n  10 MPa
2.3 2.3
12.7 10.8
ii) K s ≈ ≈ 5.5 MPa/mm at n  20 MPa ii) K s ≈ ≈ 4.7 MPa/mm at n  20 MPa
2.3 2.3

Table 16.10 Barton-Bandis modelling of the adverse effect of significantly affected as a result of the assumed reduc-
water saturation on the conducting apertures of a tion in wall strength JCS due to moisture.
joint at two different normal stresses. The incremental normal stiffnesses can easily rise by a fac-
n  10 MPa n  20 MPa tor of 10 to 100 when joints in weaker rock are ‘imaged’ at
an effective normal stress of 10 or 20 MPa, as may apply in
1. JCSo dry  70 MPa e  5.1
m e  1.6
m
a reservoir. As we also saw earlier, from Bandis normal stiff-
E  30.3
m E  16.8
m
2. JCSo wet  50 MPa e  3.9
m e  1.2
m
ness testing, Kn can rise to 103 or 104 MPa/mm if effective
E  26.5
m E  14.4
m stresses become close in magnitude to the saturation-
reduced JCS values. This is logical, since the joints in ques-
tion will be almost closed. Thus, fluid saturated joints in
moderate 13 to 15% reduction in shear stiffness from weaker rocks may have very high ratios of Kn/Ks. We will
the dry to saturated cases, in the examples in Table return to this theme of ‘almost closed’ joints in greater
16.9, in which moderate assumptions for effects of sat- detail later.
uration have been made. Concerning the anisotropic shear wave splitting and
A similar exercise for normal stiffness is more easily polarization into the fast qS1 and slower qS2 wave
performed using the BB numerical model. A cursory speeds, the former parallel and the latter perpendicular
application shows that the initial normal stiffnesses to the dominant joint or fracture sets, one may assume
(where Kni is the stress-closure gradient at very low that it is the relative low value of KS (or the high value
stress), are in the region of 28 to 32 MPa/mm with the of shear compliance ZT), rather than the high value of
above input data. Normal stiffness rises by a factor of at Kn, that is most responsible for shear wave splitting. It
least thirty at 20 MPa normal stress, on cycle #1. would also appear reasonable to assume that very tight
With the foregoing assumptions for input data (JCSo (low-permeability) joints, with high normal stiffness,
dry  70 MPa, JCSo wet  50 MPa, it is found from BB will tend to give faster qS2 wave speeds. Shear-wave
modelling that the initial (first load cycle) normal stiff- anisotropy, and attenuation anisotropy might therefore
nesses are relatively little affected by the dry: wet con- tend to be increased. Theoretical treatment in this area
dition. However, the final apertures, being so small, are of ‘compliance-anisotropy’ was given in Chapter 15.
Joint stiffness and compliance and the joint shearing mechanism 513

16.6 Mechanical over-closure,


thermal-closure, and joint
stiffness modification

Unfortunately, when joint samples are recovered from a


borehole for laboratory testing, or exposed in a tunnel
or mine adit for in situ testing, the object of our inves-
tigations is ‘cold’, in relation to its likely formation
temperature. The laboratory may also be ‘cold’ in rela-
tion to the sample’s recent environment. Several pieces
of experimental evidence from heated lab-tests, from
heated block tests and from heated ‘mine-by’ tun-
nelling experiments, suggest that the rock joint normal
stiffness, and therefore the apparent rock mass deform-
ation modulus changes, when the rock is heated in the
presumed direction of higher joint formation tempera-
tures. What appears as a measured reduction in stiffness
and modulus, is actually producing a stiffer medium.
Clearly this process is also going to affect normal com-
pliances, particularly when rock quality is high, and
Estatic is closer to Edynamic.
Both indirect circumstantial evidence, and directly
measured evidence for these phenomena, were reviewed
by Barton and Makurat, 2000, in an NGI contract
report (‘Hydro-Thermal and Mechanical Hysteresis
and Over-Closure Effects in Joints and Rock Masses
caused by Mechanical and Thermal Loading and
Unloading’). Some published results of such tests were
given by Barton et al., 1985, discussed in Barton,
2004b, and reported by Barton and Makurat, 2006.
One of the best documented cases, and also the first Figure 16.29 Idealisation of average local jointing in the 2 
of very few hydro-thermo-mechanical in situ tests in 2  2 m TerraTek/CSM heated block test in quartz
monzonite. This test gave the first well instrumented
rock mechanics, was described by Hardin et al., 1981,
in situ data on MHT coupling, suggesting that joint
and further analysed by Barton, 1982. The 8 m3 in situ
apertures closed under the effect of increased tem-
test set-up, allowed biaxial (normal) loading, or uniaxial perature, despite carefully controlled normal stresses.
loading in ‘N–S’ and ‘E–W’ directions, i.e. shearing. Note the flat-jack loading of four vertical boundaries,
Flat-jack loading on four (or two) sides of the sawn-in- and the line of borehole heaters. Hardin et al., 1981.
place block of quartz monzonite provided the desired
stress levels. There was also a line of borehole heaters
to raise the temperature by some 60°C. A simplified The flat-jack loading test cycles are shown in Figure
schematic is shown in Figure 16.29. 16.31. Maximum stress was generally 1000 psi, or
The mean joint spacing in the area of the block test is 6.9 MPa. The very first unload-load-unload-load cycles
indicated. The aspect of the test to be described here is are shown in Figure 16.32, plotted in terms of normal
the permeability test holes, drilled parallel to the long stress versus hydraulic aperture (e), where permeability
diagonal, foliation joint. Jointed core extracted from during low-pressure-gradient laminar flow was assumed
these holes is drawn in Figure 16.30, and shows the to be given by K  e2/12. The vertical axis of the figure
measured joint roughness traces, and the tilt-test failure shows the e  38
m hydraulic aperture before the
angles, giving JRCn estimates of 8.0, 8.3 and 7.9 due to block was line-drilled to make space for the boundary
the 200 to 300 mm long samples, while the mean JRCo flat-jacks. This unloading caused the hydraulic ape-
was 13. rture to increase to e  61
m. Following a minor
514 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.30 Three of the axially jointed core from the block test Figure 16.32 Hydraulic aperture (e) versus normal (applied) stress
permeability test holes, with as-recorded roughness for the first ambient temperature, biaxial load-unload
profiles, correct tilt angles at shear failure (70.1°, cycle. Note apertures before and after drilling of the
72.1° and 69.8°) and back-calculated JRCn rough- flat-jack slots. Hardin et al., 1981. (See normal stiff-
ness coefficients of 8.0, 8.3 and 7.9 at the given ness calculation, and possible normal compliance
length scales. Barton, 1982. interpretation, in text.)

Figure 16.31 The main boundary-stress-temperature test paths, for the HMT heated block test. Hardin et al., 1981.
Joint stiffness and compliance and the joint shearing mechanism 515

0–3.5–0 MPa ‘dress-rehearsal’ test cycle, the first ambi- end of the roughness and strength scales for petroleum
ent 0 to 6.9 to 0 MPa biaxial test is shown, with 1  2, reservoir rocks, and perhaps relate best to the stronger
n  6.9 MPa and  0 MPa. carbonates. The uniaxial strength of the present igneous
rock was however, ‘too high’. The JCS  c mismatch
was due to slight alteration of the rough joint walls.
16.6.1 Normal stiffness estimation

The typical increasing normal stiffness seen from 16.6.2 Thermal over-closure of
n  3.45 to 6.9 MPa, mirrors laboratory testing. If joints and some implications
one had assumed e  E, then the normal stiffness in this
increment would have been 3.45/(34.4 – 30.0)  When temperature was applied in the above block test,
103 MPa/mm  784 MPa/mm. However, the phys- unexpected over-closure of the diagonal test joint was
ical apertures can be estimated to have reduced from experienced. Figure 16.33 shows a three-parameter
144.8
m to 135.2
m, making E  9.6
m. This (coupled-process) plot of (e) versus ( n) versus tempera-
estimate is based on JRCo  13 and application of the ture (°C), and the following result.
empirical model e  E2/JRCo2.5, shown in Figure 16.7. The added joint closure means that this cooled (sam-
The interpreted normal stiffness is then about pled), unloaded joint is displaying an apparent reduction
360 MPa/mm, which is exactly in line with Kn labora- in normal stiffness. This will be estimated shortly. In the
tory test results reviewed earlier in this chapter. reality of e.g. a warm-shallow, or hot-deep petroleum
If we make the assumption that the rock quality was reservoir, joints are likely to be better interlocked than
sufficiently high for Estatic  Edynamic, this Kn value of when we unload and cool them for (initial) laboratory
360 MPa/mm would imply, by inversion, a normal testing. In fact it would be better to keep them hot
compliance ZN of about 0.0028 m/GPa, or 2.8 
1012 m/Pa. In fact the rock mass quality was very good,
as the following estimation of Q-parameters shows (see
Appendix A for ratings).

100 3 1
Q    ≈ 67 Q c  Q  c
9 1 0.5 100
200
Q c ≈ 67  ≈ 130 (‘very good’)
100

Chapter 13, Figure 13.59 relating Qc – Vp – M –


depth – porosity, suggests a static modulus of deformation
as high as 70 GPa, when taking into consideration the
favourable mine-adit-based (and applied) stress level,
equivalent to 300 m depth, and a nominal (hard rock)
1% porosity. This figure also suggests a seismic refraction
Vp close to 6 km/s for such good rock, located next to the
cross-hatched ‘hard-jointed’, ‘hard-massive’ boundary.
One could then reasonably assume that the measured Figure 16.33 Hydraulic aperture (e), versus normal stress ( n),
normal stiffness Kn  360 MPa/mm derived from the versus average rock temperature (T°C) in the per-
meability test volume of the TerraTek heated block
permeability test ( under a 3.45 to 6.9 MPa increment of
test, CSM mine, Colorado. Note aperture (e) reduc-
normal stress), was almost equivalent, in this case, to a
tions from 30.0
m to 18.3
m, to 12.9
m and
normal compliance ZN  2.8  1012 m/Pa. The roughly finally to 9.1
m as a result of temperature rise,
1 m size result would ‘plot’ immediately above the despite constant applied stress. This gives an appar-
(Hudson) diagonal line in Figure 16.23. ent reduction in the normal stiffness in this test, but
The joint characterization of the permeability test hole in the warmth of a deep petroleum reservoir, would
cores, gave JRCo  13 and JCSo  90 MPa. One could have allowed joints to remain stiffer since their for-
estimate that these values would both be at the upper mation. Barton, 1982. see Plate 11.
516 Rock quality, seismic velocity, attenuation and anisotropy

Table 16.11 Effect of temperature on joint (hydraulic) apertures enough to explain the need for several load-unload-load
at the heated block test. cycles to move beyond the strong hysteresis ‘always’
Test No. 11 12 13 16 experienced on the first load cycle, as emphasised by
Barton, 1971, Bandis, 1980 and Bandis et al., 1983, and
n (MPa) 6.9 6.9 6.9 6.9 demonstrated in Figures 16.17.
e (
m) 30.0 18.3 12.9 9.1
To illustrate the order of magnitude involved, from the
Temp°C 12° 41° 55° 74°
above coupled HMT block test, one may utilise the ambi-
Note: water viscosities corrected for T°C, before calculating (e). ent Kn value of approximately 360 MPa/mm as a starting
point. This value was established under ambient condi-
tions, with a final stress increment of 3.5 to 6.9 MPa,
before testing, and test them hot too, to get the most therefore reaching the same stress level as applied in the
relevant response, whether for simulating in situ normal heated part of the test. The constant 6.9 MPa was carefully
stiffness, or dynamic compliance. The key parameter monitored during heating of the block, to avoid thermal
affected by thermal-closure is the thermal expansion expansion-caused increases in flat-jack pressure. The suc-
coefficient. Joints must be included, and tested hot. cessive reductions of conducting aperture were from
The explanation for this phenomenon was assumed 30
m to 18.3
m to 12.9
m to 9.1
m, for tempera-
to be quite simple (Barton, 1982). Namely that the joint ture increases from 12°C to 41°C to 55°C and finally to
in question, and perhaps the huge majority of joints 74°C. The process took about 1 month: in geophysics ter-
developed in the crust, were formed at variously elevated minology about 107 Hz.
temperatures. They were thereby given a primeval ‘finger- As shown earlier, if one had assumed e  E, then the
print’ of 3D-roughness, which was influenced by all the normal stiffness in the previous ambient increment
minerals (or grains) forming the joint walls. When would have been equal to 3.45 MPa/(34.4 –
cooled, various subtle changes would occur, causing 30.0)
m  103  780 MPa/mm. But since E 
reduced fit. 14.0
m, based on JRCo  13 and the empirical model
In the case of igneous rocks, many of the earliest joint- E2  e2. JRCo2.5, the interpreted ambient normal stiff-
ing episodes would be at the cool side of the brittle- ness was about 360 MPa/mm, which was in line with
ductile transition. When such a joint, any joint, is (ambient) Kn test results reviewed earlier in this chapter.
encountered today, (e.g. at shallow depth in a mine, or The estimations of thermally-changed (static) normal
if drilled at depth and bought to the surface causing stiffnesses (and therefore potentially changed dynamic
unloading and further cooling?), the 3D roughness fin- compliances) are set out in Table 16.12, for transparency.
ger-print, though very recognisable, would be subtly The assumption is made, for purposes of calculation,
altered in its finer details. These details (very important that each thermally-induced reduction in aperture
at the micron-scale) would be a combination of: occurred at the end of the last load increment. The suc-
cessive closures are therefore allowed to have an accu-
a) a slight expansion from (anisotropic?) stress relief mulative effect on normal stiffness, as if each test started
when sampled from depth at 12°C and 3.45 MPa normal stress. In fact, since a
b) a sampling damage, but neglected for sake of strict definition of stiffness is stress increase divided by
simplicity aperture reduction, one could claim that the joint had
c) a general thermal contraction effect, but this might ‘zero’ stiffness, since no stress increment (rather a ther-
allow the equally altered finger-prints from each mal increment) was applied.
wall to still fit perfectly If this (almost reversible) thermal over-closure phe-
d) locally inhomogeneous contraction superimposed nomenon is of general application in the case of the less
on c), due to non-equality of the thermal contrac- planar jointing, it implies that when the joint in question
tion coefficients of the constituent minerals and is ‘sampled at depth’ by seismic anisotropy determina-
grains along the joint walls tion, in the familiar warmth of a petroleum reservoir, the
elevated stress and temperature will likely result in lower
Factors a) and d) are good reasons for less than ‘per- compliance, due to its previously imprinted greater stiff-
fect fit’ of the primeval joint walls in a testing situation, ness. However, the present temperature at depth is
when initially testing at ambient temperature. Factors a) unlikely to be the same as the temperature when the joint
and b) alone, without temperature considerations, are was formed.
Joint stiffness and compliance and the joint shearing mechanism 517

Table 16.12 Estimation of thermally-induced, apparent reductions of (static) normal stiffness, from the heated block test.
For Kn estimation assume each test starts at 12°C and 3.45 MPa, and ends at the specific elevated temperature.

Temperature Hydraulic Physical aperture


increment aperture (e)
m e (
m) E (
m) estimated E (
m) Kn (MPa/mm)

12 to 12°C 34.4 to 30.0 4.4 144.8 to 135.2 9.6 360


(ambient)
12 to 41°C 30.0 to 18.3 11.7 135.2 to 105.6 29.6 88
41 to 55°C 18.3 to 12.9 5.4 105.6 to 88.6 17.0 61
55 to 74°C 12.9 to 9.1 3.8 88.6 to 74.5 14.1 49

The joint will also be open to changes bought about strength in relation to 1:1 control samples, in the case
by the partial cooling caused by e.g. water flooding. The of rough tension fractures which remained ‘over-closed’
above mechanisms suggest a ‘better than expected’ aper- following removal of the pre-stress. This caused diffi-
ture and permeability enhancement as a result of water culties with over-stable modelled rock slopes, in 40,000-
flooding. Efforts to reduce the natural warming of this block tension fracture models. Barton, 2004b, discussed
water on the way down the injection well, could be this phenomenon as being due to ‘JRC at right angles’,
beneficial too, but avoidance of matrix contraction and (i.e. an effective JRC in a perpendicular sense relative to
oil-bypass (unstable fingering) might be a contrary the joint plane). This causes effective locking of asperi-
requirement. ties, if roughness is sufficient, and the stress reduction
This hydrothermomechanical HTM coupling ‘detail’ significant. When JRC is sufficiently low, such as less
(in fact a quite important experimental detail), has not than 10, possibly only less than 6, the phenomenon is
yet been absorbed into numerical modelling. It has not not expected (refer to Figure 16.26 100 mm roughness
even been coded in the Barton-Bandis joint model, profiles in this context).
mainly because data is presently limited, as the phe- The writer has experience of a very rough tension frac-
nomenon is little known, researched, or acknowledged ture, generated to make a sample for demonstrating tilt
in the rock mechanics (modelling) community. Barton testing, actually tolerating a tilt angle of 180°, i.e. going
and Makurat, 2006 have addressed the problem once from horizontal (no shear, pure normal stress from the
more, in an attempt to bring over-closure (in rock self weight of a carefully placed upper half of the block)
mechanics) into focus, as it has been in soil mechanics to 90° (vertical) to 180° (upside-down). In other words
for decades, but regarding a somewhat different mech- even the small pre-stress from self weight loading,
anism (i.e. hard, over-consolidated clays, with their amounting to about 0.02 MPa, was sufficient to give an
altered matrix properties due to unloading from higher apparent tensile strength to the fracture, due to asperity
historic effective stress levels.). inter-lock.
Barton, 1973, discussed similar experiences reported
by a colleague at Imperial College, who registered an
16.6.3 Mechanical over-closure error when preparing a direct shear test. An over-closure
episode prior to shear testing (M. de Freitas, pers. com.
There is another unresearched aspect, which may have 1970) caused a jointed sample to be too strong to shear,
application to shear wave splitting phenomena (and even though the normal stress was already reduced to the
assumed anisotropy), and that is mechanical over-closure. correct level. The sample required mechanical wedging
Rough joints or fractures that are historically loaded to to open it, when extracted from the over-loaded DST
a higher effective normal stress, and later sheared (or apparatus.
shear-wave loaded) under today’s lower effective normal
stress, (due for example to an episode of over-pressure),
will have a higher shear strength, a higher shear stiff- 16.7 Consequences of shear stress
ness, and a higher normal stiffness, than if convention- on polarization and permeability
ally tested at 1:1 n levels.
Over-consolidation ratios of 4:1 and 8:1 were shown In Chapters 14 and 15, when addressing P-wave and
by Barton, 1973, to cause several degrees higher frictional S-wave anisotropy in reservoirs, several case records were
518 Rock quality, seismic velocity, attenuation and anisotropy

reviewed that showed some angular discrepancy between


the various indicators of predominant joint strike (FMS,
oriented core), the assumed H max direction (90° from
borehole calliper-log long axis), and the principal axes of
either (or both) P-wave and S-wave anisotropy. The lat-
ter is obtained from shear-wave polarization when split-
ting into fast and slow qS1 and qS2 components. Figures
14.16, 14.35 and 15.60 are three of the cases.
In this section of Chapter 16, we will explore in much
greater detail, some of the reasons for believing that shear-
ing (as well as mineralization ‘bridging’) may be a neces-
sary mechanism for explaining hydrocarbon (especially
oil) production, from highly stressed joints or fractures.
Vertical, dominant fractures perpendicular to the min-
imum horizontal stress direction, as often sought and
sometimes ‘confirmed’, may be a more complex aspect
of geomechanics than usually assumed. Contrary to the
assumption of joints or fractures loaded by 20 to
40 MPa ‘minimum’ stress actually being ‘open’, there
Figure 16.34 A UDEC-BB model of the Terra Tek/CSM heated
may be: block test. Note the complex stress distribution even in
an ‘equal-biaxial’ loading situation, with only four
1. insufficient strength (specifically JCSconfined) for (assumed) interlocking blocks. Input data for the dif-
conducting joints in the assumed direction ferent joints were: JRC  10 or 13, JCS  90 or
2. there may be two sets of joints contributing their 120 MPa, r  25° or 28°. Chryssanthakis et al.,
seismic ‘components’ on either side of the assumed 1991.
H max direction
3. there may be conjugate shearing on one or more (approx.) loading. Normal stresses across the joints are
joint set and therefore permeability enhancement clearly in equilibrium, but the stress in the perpen-
(or maintenance). dicular direction can be dramatically different, with at
least 10 MPa difference, from one side of a joint to the
other side.
16.7.1 Stress distribution caused by One may visualise that the passage of shear waves
shearing joints, and possible through such a jointed (fractured reservoir) medium
consequences for shear wave could perceive the joints as both displacement discon-
splitting tinuities and stress discontinuities, even though the
classic Schoenberg 1980 assumption is for stress conti-
The in situ heated block test of 8 m3 volume, described nuity across the displacement discontinuity.
in the previous sub-section, was subsequently the object If this stress discontinuity complication (in the joint-
of distinct element modelling by numerical modelling parallel direction) was generally in operation, there could
colleagues at NGI in Oslo. Some of this work is then be four potential contributors to shear wave split-
reported in Chryssanthakis et al., 1991. The modelling ting, needing acceptance or rejection by the theoreticians
was a form of validation of the distinct element code who debate, conclude and publish such things:
UDEC, with the Barton-Bandis joint model as a sub-
routine, termed UDEC-BB in this non-linear form, or 1. the possible stress discontinuity (even when joints
UDEC-MC when with simpler linear Mohr Coulomb were under no shear stress)
joint modelling. An example of the complex stress dis- 2. the potential stiffness anisotropy (Kn dyn  Ks dyn,
tribution in even just a four-block (simplified) model is ZN - ZT)
shown in Figure 16.34. The flat-jacks were simulated as 3. the excess compliances/stiffnesses (the presumed
fluid-filled boundaries. Note the complex stress distri- source of (qS1 – qS2/qS1 anisotropy) with different
bution, even with equal (biaxial) 1  2  5.4 MPa contributions according to incident angle
Joint stiffness and compliance and the joint shearing mechanism 519

Figure 16.36 Multi-linear modelling of joint roughness with UDEC,


showing shear-displacement vectors of upper sample,
and the asperity-controlled principal stresses caused
by the simulated shearing. Gutierrez and Barton, 1994.

The limited shearing corresponds also to the reality


of this in situ test. There was no more than 0.25 mm
shear, despite exceeding the shear strength according to
conventional (Mohr) stress transformation of stresses.
This is because the conventional stress transformation
Figure 16.35 Equal biaxial ( 1  5.4 MPa, 2  5.4 MPa) load- equations, for converting principal stresses into shear
ing effect on the maximum values of the physical (E) and normal stress components, fail to take account of
and hydraulic (e) apertures. Top: 142
m, 35
m, the contribution of dilation to what are actually non-
Bottom: with dilation-limited shearing, E  169
m, coaxial stresses and displacements. The mobilized dilation
e  49
m. Chryssanthakis et al. 1991.
angle needs to be added into the cos 2  , sin 2  terms
of the classic transformation equations. Barton, 1986.
4. the presence of a variable thickness film of water If we move now to a more direct simulation of joint
between the opposed, contacting sides of non- shearing, in a simulated DST direct shear test, as
planar slightly sheared joints. opposed to a uniaxial/biaxial test, we avoid the problem
of the stress transformation error, as the principal stresses
If there has been pre-shear (in a ‘critical crust’ scen- are applied parallel and perpendicular to the joint plane.
ario), then presumably the parts of the joints that were Gutierrez in Gutierrez and Barton, 1994, used the
dilated and fluid filled would have changed local stiff- discrete element code UDEC to simulate the roughness
ness in relation to the rock-to-rock stress-transferring of rock joints in a direct, unusually realistic manner, by
parts, thereby contributing differently to attenuation, constructing multi-linear sectors to represent some of
P-, and S-wave anisotropy. the measured 100 mm long roughness profiles of
In Figure 16.35, the modelling of physical (E) and Barton and Choubey, 1977 (from Figure 16.26). One
hydraulic (e) apertures is shown, based on the assumed of the UDEC models is shown in Figure 16.36. Note
JRC  10 or 13, JCS  90 or 120 MPa, r  25° or the dilation-related inclination of the deformation vec-
28° input data. The upper pair of (E) and (e) models are tors for the ‘upper’ sample, the bottom being fixed. The
under equal biaxial stress, while the lower pair are under principal stresses applied were parallel and perpendicu-
N–S ( 1 only) loading. Slight non-uniformity of aper- lar and there was simultaneous flow of fluid simulated
ture is shown, though shearing was very limited, for from left to right, along the joint void. This particular
important reasons which will be discussed below, and joint simulation had JRCo  5. The constant normal
treated later. stress was only 1 MPa.
The maximum joint apertures shown in Figure 16.35 It is very easy to note the entirely different principal
were as follows: stress vectors above and below the shearing joint. The
1  2  5.4 MPa e  35
m only locations where there are more similarities in mag-
E  142
m nitudes, but not directions, are the concentrated regions
1  5.4 MPa, 2  0 e  49
m of stress transfer at the two major asperities, which resem-
E  169
m ble the more exaggerated case of a sheared undulating
520 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.37 Interlocked and sheared joints in ‘wavy’ columnar basalt, demonstrating the role of asperities and dilation on aperture distri-
bution. Columbia River Basalts, Washington State, USA. See Plate 12.

joint in basalt, from the Columbia River basalt sequence stress ( n), and the maximum tensile stress was 13 times
in the Western USA, shown in Figure 16.37. higher than n. Apertures were mostly ‘closed’, or
On the basis of such rock stress related phenomena, ‘opened’ to about 2 mm. (There appears to be a ‘mm for
one may then pose the following questions to geo- m’ misprint in diagram d, where 103 m and 106 m
physics theoreticians. In the case illustrated by the shear- should be the two approximate extremes).
ing joint model of Gutierrez, will shear wave splitting Finally, to round off this present discussion of the pos-
occur more readily where the principal stresses ‘plunge’ into sible consequences of an actual local stress-discontinuity at
another direction (as at the stress-transferring asperities), joint surfaces, especially when under shear stress, one may
or will shear wave splitting be able to occur across the ␴1 consider the inter-bedded bituminous shale and dolomite
stress discontinuity, as generally seen in Figure 16.36, due ‘Kimmeridge clay’ source rocks depicted in Figure 16.39a
to the general presence of a ‘fluid-filled’ lense between these and b. The views are taken looking upwards from the base
same locations? of the 10 m high cliffs at Kimmeridge Bay in Dorset.
A painstaking ‘diagnostic’ of the simulated shearing Although a certain consistency can be noted in the
joint was performed by Gutierrez. This is shown in orientation of the two (actually up to three) sub-vertical
Figure 16.38. In the diagrams we see the following dis- joint sets in the harder dolomite beds, it is easy to imag-
tributed phenomena, along the 100 mm long simulated ine the relative complexity of the joint stress distribu-
rock joint: tions, in view of the stress discontinuity trend discussed
above. Interpretation of shear wave splitting in such a
a) individual contact lengths (mm) for 0, 1 and 2 mm
fractured reservoir environment, perhaps with non-
shear
aligned horizontal stress anisotropy, can be imagined as
b) individual contact angles for 0, 1 and 2 mm shear
somewhat demanding.
c) stress to strength ( n/JCS) ratios for 0, 1 and 2 mm
shear
d) contact apertures for 0, 1 an)d 2 mm shear 16.7.2 The strength-deformation
components of jointed rock
There is probably a correlation of high stress to masses
strength ratios where there are small contact lengths and
high contact angles. The highest (compressive) stress The potential for shear-wave splitting as a means of mon-
concentration was 38 times the applied (average) normal itoring temporal changes in the geometry of the aligned,
Joint stiffness and compliance and the joint shearing mechanism 521

Figure 16.38 Numerical diagnostic of the sheared UDEC joint, showing contact lengths, contact angles, contact apertures and stress ratios,
with 0 or 0.5 mm, 1.0 mm and 2 mm of shearing. Gutierrez and Barton, 1994.

Figure 16.39 Vertical view through the inter-bedded bituminous shales and dolomite bed, in Kimmeridge Bay, showing a) complex,
b) ordered sub-vertical joint patterns. Dorset coast, S. England. See Plate 13.
522 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.40 Depiction of the effect of block size on the shear strength components of joints, from Bandis 1980, and Bandis et al., 1981.

fluid-filled features responsible for the splitting, can be Evidence for this was obtained from acoustic emission
further visualised by reference to cases where shear monitoring, which showed migration of micro-shearing
deformation occurs in the course of production from a to successively greater depths, when attempting to estab-
petroleum reservoir. Due to the usual non-planarity of lish hydraulic connection between wells, for extraction of
joints (despite their obvious depiction as straight lines or thermal energy. A schematic of such a case was illustrated
planes in the literature), there will usually be non- in Chapter 15, Figure 15.2a, and temporal effects on
linear normal stress-closure-opening behaviour, and non- shear-wave anisotropy were documented.
linear shear-dilation behaviour, Some of the details of The representative block size, given by the number of
joint behaviour need to be explored further, if stress joint sets and their spacing and orientation, plus the
anisotropy causes shearing during production, and if seis- surface character and length of the joints, combined
mic monitoring is to be used for diagnosing the effects with the effective normal stress levels acting on each set,
of new production measures. collectively contribute to the possible response of a rock
There is the potential for shearing when producing mass to shearing, when under the influence of an
from conjugately jointed reservoirs when a vertical anisotropic, and non-aligned principal stress field.
principal stress roughly intersects the two steeply dip- Figure 16.40, from Bandis, 1980, shows the dramatic
ping sets of conjugate joints in the reservoir (cf. Ekofisk, potential effect of the individual block size, on shear
as described in Barton et al., 1986). This case will be strength, and displacement to peak, especially when
discussed later in this chapter. joint roughness is significant.
The potential for shear also exists when water flood- Barton and Choubey, 1977, had anticipated the effect
ing into a jointed reservoir with non-aligned, anisotropic of block size on displacement to peak (peak), as shown
principal stress, for enhancing production. A related conceptually in Figure 16.41. Despite identical JRCn
case is the injection of cold water into a geothermal for joints AA and BB, the altered block size was
reservoir, to establish hydraulic connection to wells on expected to reduce peak. Clearly the bulk modulus of
the far side of the reservoir. the second case would nevertheless be lower than for
In the first case a gradual but major reduction in pore the massive rock mass, as there are so many more joints
pressure (at least 20 to 24 MPa, prior to water-flooding) involved in the deformation process.
had caused an equivalent effective stress increase, fol- Confusingly perhaps, it may be the ‘better quality’
lowing some 15 years of production. The porous chalk rock masses, with wider joint spacing, and with greater
matrix apparently contracted sufficiently to make space joint continuity, that could be most prone to (previous)
for down-dip shearing, despite one-dimensional com- shearing deformation along the joint sets, when under
paction, (see Figure 15.25). the influence of a non-aligned, anisotropic, (differential)
In the second case the injection caused an increase in stress field, caused by some historic tectonic adjustments,
pore pressure, and sufficient reduction in effective normal of which there is no shortage.
stress for the unequal principal stresses (70 and 30 MPa, A heavily fractured rock mass, though obviously with
Pine and Bachelor, 1984) to cause slight shearing. lower (bulk) modulus, may resist significant shearing
Joint stiffness and compliance and the joint shearing mechanism 523

for their first (photogrammetry-measured) shearing


events of 21.6, 25.1 and 26.7 when tested in biaxial
shear, i.e. shear strength rising as block size reduced.
The mass ‘Poisson ratios’ estimated from these ‘static’
biaxial shear tests are seen to rise dramatically as differen-
tial stress rises. Values exceed the continuum ‘limit’ of
0.5, due to the influence of micro-shearing along numer-
ous fracture surfaces. Presumably, in the context of a
sheared reservoir situation, such could be detected by the
seismic anisotropy response. Shear wave anisotropy would
be enhanced most by the more easily shearing (and there-
fore more conducting) primary set of fractures. (Note that
these would not be parallel to the customary H max
direction).
In these physical tension fracture models, the first
parallel set of fractures developed on the double-guillotine
table, were continuous, and remained continuous, fol-
lowing the development of the secondary, intersecting
set, which due to up-stepping and down-stepping when
crossing the pre-existing fractures, would impart an
anisotropic strength and stiffness to the assembly. It is
reasonable to believe that in the context of conjugate
joint sets, this dominance of a primary set is common.
It was certainly the case in the jointed chalk of the
Ekofisk field, as reviewed in Chapter 15, Figure 15.7.

Figure 16.41 Scale effect on  (peak) caused by increasing block


size. Barton and Choubey, 1977. The scaling of peak 16.7.3 Permeability linked to joint
is one of the components of Ks that causes low stiff- shearing
ness as block size increases.
In a deep petroleum reservoir, and strictly from a geo-
mechanics viewpoint, one can hardly imagine that
along individual joints, due to the higher collective ‘open’ joints can be found at typical 2 to 4 km reservoir
shear resistance of smaller block sizes. When the block depths, unless there are one or more of the following
sizes are small, blocks can rotate slightly, thereby sam- conditions: mineral ‘bridging’ holding parts of fractures
pling the steeper joint-surface asperities that contribute open, high-strength rock and rough-surfaced joints, or
to higher strength. (Barton and Bandis, 1982). joint shearing (possibly where there is less roughness).
Some physical modelling evidence of the influence of This viewpoint is held because of a probably too high
block size is shown schematically in Figure 16.42a, effective minimum stress ( h min) of perhaps 20 to
from Barton and Hansteen, 1979, as also analysed fur- 40 MPa magnitude, in relation to probable similar
ther by Bandis et al., 1981. The (differential) stress- magnitudes of JCSn in the case of the common weaker
deformation behaviour of the three different models is reservoir rocks. Shearing, as in weak Ekofisk chalk, and
shown in Figure 16.42b. Back-analysis showed that the perhaps at a larger number of reservoirs than the petro-
elemental tension fractures, common to each model, leum industry suspects, is linked of course to the poten-
whether forming 250, 1000, or 4000 interlocked tial for dilation, and therefore potential permeability
blocks, had a ‘JRCn’ value averaging 20, from numer- maintenance (in the face of the high normal stresses),
ous ‘large-scale’ direct shear tests. The tension-fractured and perhaps even permeability enhancement.
assemblies, made with a dynamic ‘double-guillotine’ Example values of the three basic components of joint
device (Barton, 1972), each fabricated of identical rock- or fracture shear strength and stiffness (JRC  roughness,
like, brittle material, had back-calculated JRC values JCS  joint wall compressive strength, r  residual
524 Rock quality, seismic velocity, attenuation and anisotropy

Peak strength
of jointed mass

(a)

(b)

Figure 16.42 a) A comparison of the shear strength of interlocked assemblies of tension-fractured, idealised, model rock masses, with 250,
1000 or 4000 blocks. b) The respective differential stress-versus-strain curves in the 1 and 2 directions for the same models.
Note entirely different behaviour of the most heavily fractured model with 4000 blocks, due to greater freedom for (micro)
block-rotation. Barton and Hansteen, 1979.

friction angle) are shown contributing to peak shear amounts of predicted dilation, with the given input
strength in Figure 16.43. Estimated peak dilation angles are data assumptions, are shown in the middle diagrams. In
also shown on each strength envelope, at appropriate nor- the lowest of the three sets of predicted behaviour
mal stress levels. Clearly as shearing continues beyond curves, the theoretical maximum change in permeabil-
peak, rates and angles of dilation will steadily reduce. ity with the shear-induced dilation is shown. Here,
However the joint (minor fault?) aperture has probably gouge production (and partial joint blocking) is
accumulated some vital void space for fluid conduction, ignored. Note the assumption of an initial hydraulic
minus areas of contact with crushed material and finer aperture (eo) of 25
m in each case.
gouge causing local blockage, as we saw in Figure 16.13. The dilation curves show how the initial physical
Barton-Bandis modelling of individual joint responses aperture (Eo) changes. The sum of Eo  E (E1) is
as either block size or normal stress changes are shown in converted to eo  e (e1) to estimate permeability
Figure 16.44, from Barton et al., 1985. The relative from the assumed cubic law K  e2/12.
Joint stiffness and compliance and the joint shearing mechanism 525

fracture related rather than stress-aligned microcrack


related (EDA).
In Figure 15.60, we could see the various sources of
orientation data from Lynn et al., 1999. There was
clearly great consistency in the direction of the N20° to
35°W joint set, which was presumed by Maultzsch et al.,
2003, Liu et al., 2003a and Liu et al., 2003b to be the
source of the qS1 polarization. Yet diagram b) in Figure
15.60 showed ‘VSP S1’ at about N43° to 45°W, some
10° to 20° different from the fracture orientation
obtained from core data, yet almost parallel to the
gilsonite dykes.
One may speculate that a possible explanation for the
potential discrepancy between qS1 polarization and the
fracture orientation could be due to non-planarity of
the fractures combined with slight shearing. It could also
be due to unavoidable, minor sources of error, in data
acquisition, as suggested by Lynn (pers. comm. 2005).
The Chapman modelling reviewed in Chapter 15,
had suggested 6 m diameter fractures at a frequency of
Figure 16.43 The peak shear strength equation for in situ scale rock only one per 9 to 10 m cube of rock mass, i.e. implying
joints, and example strength envelopes. The sub- quite large block sizes, but this would be dependent on
scripts on JRC: (JRCn) and on JCS: (JCSn) imply the spacing of the N60° to 80°E set.
large scale, block-size-dependent values in general. In
With large block size and non-coaxial major princi-
the two uppermost strength envelopes shown, input
pal stress one could have a situation in the reservoir
values are more typical of laboratory samples, where
roughness tends to dominate behaviour. The lower
(especially if fluid pressures had been high) that, in
envelope might apply to a minor fault, or to a weak, principle, could resemble the shear-dilation mechanism
or weathered rock joint. shown in Figure 16.45. These are rough tension frac-
tures that have a prototype scale of many metres length
(due to up-scaling of a weak brittle model material).
16.7.4 Reservoir seismic case The measured profiles of roughness have been sheared
records with possible following the shear-dilation (X, Y coordinates) meas-
shearing ured during direct shear tests at a) low normal stress
(top: n 2) and b) at high normal stress (bottom: n 6).
Seeking more detail about the fracture directions and Two different degrees of pre-peak and post-peak shear
stress direction at the Bluebell-Altamont Field, one may and dilation are shown.
note from Lynn et al., 1999, that the borehole elonga- The black (overlapped) contacting areas that transfer
tion data indicated a range of possible major stress orien- shear and normal stress have, on average, a significantly
tations from N20° to 50°W. Fracturing could different orientation to those (white) areas that are dilated,
presumably have influenced this result. Three methods and which in a ‘slightly-sheared’ reservoir situation
of fracture orientation (outcrops, FMS log and core would contain most fluid. Could it be that the rotated
samples) indicated a spread between N20° to 35°W, mean direction of these ‘fluid lenses’, obviously distrib-
with the second set less well delineated at about N60° uted in three dimensions, could be more responsible for
to 80°E. the shear wave polarization than the average joint plane,
The Maultzsch et al., 2003, Liu et al., 2003a and Liu or would it be the rock contact areas?
et al., 2003b analyses of shear wave anisotropy, reviewed These large, rougher than normal tension fractures
in Chapter 15, actually suggest polarization more show a 10° to 20° effective rotation of the lenticular
related to the major stress direction than the dominant apertures in relation to the mean fracture planes. They
fracture direction. However, the variation of anisotropy might represent the areas where squirt attenuation was
at lower frequencies suggests that their results must be least active, due to higher permeability.
526 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.44 Barton-Bandis models of shear strength-displacement-dilation-permeability coupling. Curves generated by Bakhtar with a
programmable calculator. (Barton et al., 1985.) The permeability estimate is the theoretical, dilation-produced maximum,
devoid of considerations of gouge production. The latter requires a modified treatment of roughness for (E) to (e) conversion,
using JRCmobilized, as shown by Olsson and Barton, 2001. (See Figure 16.9).

The oppositely rotated rock-to-rock contact areas 16.7.5 The apertures expected of
with their extremely high stress, are less likely planes for highly stressed ‘open’ joints
reflection. So if the lenses were fluid-filled, an apparent
rotation between the reflectiveness to S-wave and the As introduction to the question of ‘open’ fractures and
squirt attenuation source areas might be the result, in their orientation, we refer to another case record of reser-
relation to the mean plane of reference, some 10°–20° voir seismic anisotropy investigation, this time involving
between the two. a large (43 square mile) 3D full-azimuth, full-offset
Previous figures in this chapter showed that shear P-wave survey in a fractured carbonate gas field in Texas,
strengths, shear displacements, dilations and shear stiff- as described by Lynn et al., 2000. The deep target forma-
nesses, were each affected adversely by increasing block tions were carbonate and chalk formations at 14,000 to
size. (Barton and Bandis, 1982). Due to reductions of 15,500 ft (about 4,280 to 4,730 m). The faster P-wave
effective roughness JRC and wall strength JCS, and velocity direction was parallel to the strike of the local
increased displacement to peak, each caused by larger structure, and interpreted as also parallel to the local max-
block size, the in situ (pseudo-static) joint shear stiff- imum horizontal stress direction. As the authors state:
ness even at 10 to 20 MPa effective normal stress, may ‘The interpreted open fractures are those approximately par-
be close to 1 MPa/mm for the case of say 2 m in situ allel to the local maximum horizontal stress, and are con-
block sizes. Such a low pseudo-static stiffness is two or sidered to be the fluid-flow pathways or permeable conduits’.
more orders of magnitude softer than the ‘Hudson com- Clearly, at four-and-a-half km depth, one is far
pliance’ diagonal (Figure 16.23). Is this also significant? removed from the near-surface, where H max oriented
Joint stiffness and compliance and the joint shearing mechanism 527

Figure 16.46 A collection of large scale stress-permeability tests on


joints in hard crystalline rocks, both at ambient
and elevated temperature, showing ‘linear’ log-log-
permeability-stress trends, with various gradients.
Barton, 1982. (Pratt et al., 1977, Witherspoon et al.,
1977, Iwai, 1976)

angle to the major stress. Such a scenario for the ‘open’


joints has been emphasised and convincingly demon-
strated in deep well analysis, by Colleen Barton, Zoback
and Moos, 1995, and this and later studies by Zoback
and his colleagues will be reviewed later in this chapter.
Before considering some convincing evidence for the
possible/probable non-alignment of ‘open’ sub-vertical
Figure 16.45 Reconstructed shear-and-dilation events for model fractures with the major principal stress direction ( H max),
tension fractures with the illustrated roughness profiles,
from these and other authors, one must consider the
cut out as plastic ‘replicas’. Note the ‘opposite’ rotations
case of fractures that are orientated in this classically
of the (potentially) fluid-bearing lenses, which are
down-dipping to the right, and the up-slope to the
assumed ‘mode 1’ direction. In such cases, the mini-
right rock-to-rock contact areas which are of ‘double’ mum effective stresses may tend to keep the ‘open’ frac-
thickness therefore signifying the production of tures actually more closed than open, unless the rock is
crushed particles. The lenses will therefore have rather strong, and/or the fractures are ‘bridged’ by min-
debris/gouge at their extremities. Barton, 1973a. eral cementation.
Even for joints in hard crystalline rocks, coupled
joints are certainly very typical. Such cases were (MH) stress-permeability tests of joints at mostly in situ
reviewed in Part I of this book, and were the source of scale (1 m diameter jointed cylinders, and 2 to 8 m3
marked azimuthal P-wave anisotropy. In contrast to the jointed in situ blocks), demonstrate a marked reduction
near-surface, the adverse effect of many tens of MPa in permeability with normal stress. Figure 16.46 from
minimum effective stress would have suggested to a rock Barton, 1982, shows ‘linearity’ on a log-log plot of per-
mechanic that the more ‘open’ joints should be at an meability versus normal stress. One earlier ambient
528 Rock quality, seismic velocity, attenuation and anisotropy

The laboratory normal stress-conducting aperture


data shown in Figure 16.48 was assembled by Makurat
(pers. comm. 2006), from numerous tests performed
in the CSFT apparatus developed by Makurat at
NGI (Makurat et al., 1990, Makurat, 1996, Makurat
2006). As is readily observed, there are a large number of
test data showing extremely small apertures, often well
below 10
m, when even moderate normal stress levels
in relation to JCS values were applied. The data for tuff
and ignimbrite joints from the UK Nirex Ltd planned
Sellafield Rock Characterization Facility, were derived
from tests at up to 30 MPa normal stress. Although this
is equivalent to effective stress levels at about 1.5 km
depth in a petroleum reservoir, the ratio of applied
n/JCS was less than 0.15, due to the high strength of
the tuff. The fact that the joints were quite planar, with
JRCo mostly from 3 to 6, was an important contributor
to their small conducting (and mechanical) apertures
under stress.
It was noted by Makurat that tests on weak reservoir
rocks like chalk and shale, often showed higher con-
ducting apertures than expected, in relation to quite
high ratios of n/JCS, such as 0.7, 1.4 and even 2.0.
This was interpreted as due to sampling damage, as it is
Figure 16.47 Log-linear relation showing the effect of normal
easy to ‘lose’ material (prior to reaching the laboratory
stress on the closure of physical aperture for various environment), due to recovery by drilling and subse-
rock joints, in mostly sedimentary rocks, as tested by quent transport. Loose grains and cracked pieces can be
Bandis, 1980. more easily lost when strength is in the 2 to 5 MPa
range, as for several of the JCS values of joints in the
chalk, sandstone and shale.
The hoped for ‘open’ joints in a fractured reservoir can
temperature and one heated block test (described in hardly be expected to be significantly open, unless the
Figure 16.29 to 16.33), constitute the main body of joints or fractures are rather rough, or are ‘bridged’ (but
data assembled in this figure. There was experimental not sealed) by hard minerals, thereby resisting complete
evidence from these tests of even smaller conducting closure during hydrocarbon production. Of course if the
apertures when temperature was elevated. This would fluid is over-pressured and is very close to the minimum
add to the downward trend of permeability, with total rock stress, so that the effective stress ( h min – Pf ) is
increased depth. Hydraulic apertures in the lower right small, then closure again would be limited until much
corner of the figure are down to a few microns in size. later in production.
It seems unlikely that such would be considered ‘open’ If we consider 3 km depth, and a fairly typical min-
in the context of oil producing reservoirs. imum total stress ( h min) magnitude of 40 to 50 MPa
The physical closures of joint apertures in (mostly) (compared to an assumed density-based total vertical
sedimentary rocks, as a result of applied normal stress, stress ( v) of 60 to 70 MPa), there may be a hydrostatic-
were shown by Bandis, 1980 to decline linearly against based pressure of about 30 MPa for the 3 km deep reser-
the log of normal stress. This relevant result is shown in voir. It is then clear that the effective normal stress of
Figure 16.47. The large closures seen in the weaker 10 to 20 MPa will be holding joints parallel to H max
rocks like siltstone, would be expected to nearly ‘close’ nearly closed, since they are undisturbed, and have no
such joints, and with conversion from physical aperture way of losing material, as above.
(E) to hydraulic aperture (e), small residual permeabil- We can deviate briefly to the micro-world of elliptical
ities would be expected. cracks and consider that the theoretical closure pressure
Joint stiffness and compliance and the joint shearing mechanism 529

(a)

(b)

Figure 16.48 a) Normal stress-conducting aperture behaviour from CSFT tests in natural joints in tuff, ignimbrite, granite, sandstones, shale
and chalk. b) Mean behaviour for each rock type, using power-law extrapolation. Makurat, 1996, Makurat pers. comm. 2006.

P, for the case of a stiffer elliptical crack of aspect ratio in which Ks  bulk modulus of solid material (e.g.
(), under hydrostatic stress (Walsh, 1965) is: about 40 GPa for clean sandstones), and   Poisson’s
ratio of solid material (0.17 for clean sandstones). The
closure pressure for a microcrack of aspect ratio 103 is
1  2 (16.10) then approximately 40 MPa. Using this theory, King
P  3K s  
2 (1   2 ) and Marsden, 2002, had assumed that microcracks with
530 Rock quality, seismic velocity, attenuation and anisotropy

(c)

(d)

Figure 16.48 c) Normalised stress-to-strength ratio ( n/JCS) versus conducting aperture plots, which emphasises the different rock strength
magnitudes. d) Power-law fit to mean data for a given rock type, with extrapolation to higher stress-to-strength levels.
Makurat, 1996, Makurat pers. comm. 2006.

smaller aspect ratios would not be open at applied effect- break down, as rock joints with some wall roughness
ive stresses greater than 40 MPa. appear not to close completely (Bandis et al., 1983).
If up-scaling is performed to much longer cracks and The normal stiffness merely becomes extremely high,
discontinuous joints, with aspect ratios of e.g. 105 or but complete closure is virtually prevented by asperity
less, the above theory for ‘elliptical’ cracks will presumably roughness, like a joint roughness coefficient JRCo that
Joint stiffness and compliance and the joint shearing mechanism 531

Figure 16.49 BB normal closure-permeability and shear dilation-permeability modelling. Case A ‘sandstone’ with JRC  5, JCS  25 MPa,
r  27° and c  30 MPa. Note extremely small apertures and low permeability in the left-side plots ( normal closure) and
the marked dilation and rapid permeability enhancement in the right-side plots ( shear).

is acting at right-angles. However, we would be reaching would shear wave splitting and recognisable fast and
micron size (e.g. 1 to 5
m or less) in such cases, unless slow wave speed anisotropy have been registered by the
previous shearing had occurred. geophysicists?
It would appear that the ‘conventional wisdom’ of
the most conducting joints being parallel or sub-paral-
lel to the major stress direction, which is clearly proved 16.7.6 Modelling apertures with
in the near-surface, by means of three-dimensional per- the BB model
meability testing (e.g. Quadros et al., 1999), needs to
be re-assessed in the context of petroleum reservoirs, in In order to investigate this further, the stress-closure
weaker sedimentary rocks at great depth. Will a joint ( n – E) and stress-permeability ( n – K) cycles
set with apertures of a few microns or less, really satisfy resulting from Barton-Bandis modelling of three hypo-
the description ‘open joints’, and will they cause shear thetical, but realistic joint characteristics were investi-
wave splitting when they have such a high area in con- gated. The input data assumptions are listed in Table
tact as is necessary for h minimum of 10, 20 or 16.13, and Figures 16.49 to 16.51 and Table 16.14
30 MPa to be transferred across their relatively weak show selected results of this parameter study.
joint walls? If the ‘open’ apertures are actually mainly The first two of these BB modelling figures shows
‘tape-worm-like’ channels, within ‘closed’ joint planes, that the modelling of both cases originates from an initial
532 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.50 BB normal closure-permeability and shear dilation-permeability modelling. Case B ‘carbonate’ (low roughness) with JRC  5,
JCS  50 MPa, r  33° and c  60 MPa. Note extremely small apertures and low permeability in the left-side plots ( normal
closure) and the marked dilation and rapid permeability enhancement in the right-side plots ( shear).

Table 16.13 Input data assumptions for investigating the physical


and conducting apertures likely to be available with an
unstressed aperture (Eo) of 0.14 mm (or 140
m). This
‘open joints parallel to H max assumption’. Unless joints was estimated from the following empirical equation
are rough, and the rock is hard, this orientation proves (Barton, 1982) based on Bandis, 1980, tests of numer-
unlikely to give the assumed ‘open’ joints at reservoir ous natural joints.
effective stress levels. Note subscripts JRCo and JCSo
JRCo  
representing lab-scale parameters, which are assumed to
E o (mm) ≈  0.2 c  0.1 (16.11)
be most representative for normal closure modelling.  
5 JCSo 
Case A JRCo  5 (medium rough, slightly undulating)
JCSo  25 MPa (medium weak rock e.g. sandstone) This equation makes allowance for the experimental
c  30 MPa r  27° result that initial apertures are larger when the small
Case B JRCo  5 (medium rough, slightly undulating) scale joint roughness coefficient (JRCo) is greater than
JCSo  50 MPa (medium strength carbonate rock) about 5, as in Figure 16.51 which has JRCo  10. The
c  60 MPa r  33° ratio of c/JCSo allows for the increased aperture when
Case C JRCo  10 (rough, slightly undulating)
there is weathering when JCSo  c. Clay-fillings or
JCSo  50 MPa (medium strength carbonate rock)
clay-coatings, as tackled in the rock quality Q-system,
c  60 MPa r  33°
are not considered in Barton-Bandis modelling.
Joint stiffness and compliance and the joint shearing mechanism 533

Figure 16.51 BB normal closure-permeability and shear dilation-permeability modelling. Case C ‘carbonate’ (high roughness) with JRC  10,
JCS  50 MPa, r  33° and c  60 MPa. Note improved apertures and permeability in the left-side plots ( normal clo-
sure) and the stronger dilation and rapid permeability enhancement in the right-side plots ( shear).

Table 16.14 Estimates of physical (E) and conducting (e) apertures,


for two assumed reservoir rock joint scenarios, with low the third and fourth cycles of consolidation (actually
and moderate strength rock, and low roughness. ‘reloading’, as in the Bandis experiments, to reach an
Predicted apertures are too small to be considered almost stable assumed ‘in situ’ aperture).
‘open’, with this limited roughness. 3  3rd cycle, The effective normal stresses applied were to simulate
4  4th cycle of loading. H min with an ‘open joints parallel to H max assump-
Case A Case B tion’, as referred to by many geophysicists with presum-
ably the agreement of the field reservoir engineers. Hard
Input data JRCo  5 JRCo  5
rock and rough joints (or ‘bridging’ with hard minerals)
JCSo  25 MPa JCSo  50 MPa
seem to be needed for this orientation to satisfy the ‘open’
At n  10 MPa E3  0.90
m E3  2.26
m
n  20 MPa E4  0.37
m E4  1.03
m
joints requirement, judging by the very small apertures
At n  10 MPa e3  0.01
m e3  0.09
m predicted for these two least-rough-joints cases.
n  20 MPa e4  0.01
m e4  0.02
m One set of conducting fractures will give the rock
mass an (anisotropic) mass permeability equal to Km
Both models with the least rough joints were consol- (following Louis, 1967):
idated three times to 20 MPa, then to 10 MPa on the
4th cycle, as if there was a certain overpressure. They e
Km  Kr  K j (16.12)
show extremely small minimum apertures (E3, E4) on L
534 Rock quality, seismic velocity, attenuation and anisotropy

where Kr  permeability of rock matrix 16.7.7 Open joints caused by


Kj  joint permeability anisotropic stress, low shear
strength, dilation
e  average hydraulic aperture
L  average spacing of conducting joints In contrast to the normal closure modelling of aper-
tures discussed above, the shear-dilation modelling that
is also shown in Figures 16.49 to 16.51, indicates that
Since in the highly stressed ( n  20 or 10 MPa) exam- significant apertures will be developed after some few
ples (cases A and B), e3 and e4 are in the range 0.01 to mm of shearing in all three cases, even with the weaker
0.07
m, it is clear that both the Kj  e2/12 term, and the of the ‘reservoir’ rocks. So repeated shearing in a ‘crit-
e/L term will give minute contributions to the bulk per- ical crust’ scenario (as interpreted by Townend and
meability Km, and one will depend on the matrix perme- Zoback, 2000, and others who support the ‘geome-
ability Kr (and porosity), in each of these two cases. On chanical’ school of thought), could easily develop the
the other hand when JRC  10 (as in Figure 16.51, the apertures needed to make the above case A, and case B
apertures and permeabilities are considerably larger, and ‘hydrocarbon reservoirs’ productive, with truly ‘open’
‘open’ joints in the direction of H maximum can then be sheared joints and less need for a permeable matrix,
imagined. than seen in the normal stress-closure examples.
There is however, a dilemma exposed by the JCSo In the case of shearing under an arbitrary effective nor-
assumptions in Tables 16.13 and 16.14, concerning mal stress level, the peak shear strengths ( ) can be esti-
which magnitude of the confined joint wall strength JCSo mated from the following equations, which were
one should actually select for reservoir joint modelling. presented earlier. Block-size induced scale effects, redu-
Clearly a positive effect of confinement on strength (rep- cing JRCo and JCSo, are seen to reduce peak shear
resented by the expanding Mohr circle diameter 1 – 3), strength. They will also affect the subsequent shear resist-
should help to increase apertures somewhat, but would ance reductions, towards residual strength, the latter tak-
likely be insufficient to give ‘open’ character in the case of ing perhaps 1 meter, or more, of shear displacement.
the less rough joints in the weaker, porous reservoir rocks.
The problem for a weak reservoir rock, like porous   JCS  
 n JRCn  log   n   r 
sandstone or chalk, is that very high confinement, as   n  
when at several kilometres depth, may no longer be pos-
itive for strength development, due to the onset of plas-  L 0.02 JRCo
JRCn ≈ JRCo  n 
tic deformation and eventual pore-collapse trends. The  L o 
Mohr strength envelopes tend to curve rapidly towards  L 0.03 JRCo
a maximum shear strength with increased confinement JCSn ≈ JCSo  n 
(i.e. have strongly non-linear internal friction angles),  L o 
and therefore reach a maximum value of 1 – 3 at the
point where the Mohr envelope becomes horizontal, It will be noted that in contrast to the normal closure
usually corresponding to a ‘critical state’ line defined by modelling of the three cases A, B and C in the previous
a gradient 1  3 3 suggested by Barton, 1976. This section, there is here a necessity for r estimation, since
will be shown later. shearing is to be modelled. In Table 16.13, r  27°
At higher confinement a ‘cap’ or declining trend occurs. was assumed for the case A, weak ‘sandstone’, and
Loss of porosity or reduction in void ratio for a rock like r  33° for cases B and C, representing stronger ‘car-
porous chalk, plus a reducing rate of gain in strength, bonates’ with two different joint roughness magnitudes.
occurs at successively lower effective mean stresses as the In situ block sizes Ln of 1.0 m were assumed here in each
porosity rises from e.g. 20 to 30 to 40% and more. In con- case, in order to demonstrate the orders of magnitude of
trast, hard rocks, and especially very hard rocks, show great possible strength reduction, due to reductions of JRCo
benefit of confining stresses equivalent to many kilometres and JCSo according to the scaling equations of Bandis
depth, by not crossing the ‘critical state’ line until hun- et al., 1981. The experimental form of the JRCo and JCSo
dreds of MPa confinement, or up to tens of kilometres of reduction with block size is illustrated in Figure 16.52.
equivalent depth in exceptional cases. These aspects will be Note that the dilation curves shown in Figures 16.49
discussed further in the next section. to 16.51, were non-conservatively linked directly to
Joint stiffness and compliance and the joint shearing mechanism 535

Table 16.15 Estimation of shear strength reduction due to block


size, for the three reservoir rock scenarios. Assumed
mean block size  1.0 m. (Note the friction
coefficient format: useful for comparison with data
from the ‘critical shearing crust’ section that follows).

Case A (assume Case B (assume Case C (assume


n  10 MPa) n  10 MPa) n  10 MPa)

JRCo  5 JRCo  5 JRCo  10


JRCn  4.0 JRCn  4.0 JRCn  6.3
JCSo  25 MPa JCSo  50 MPa JCSo  50 MPa
JCSn  18 MPa JCSn  35 MPa JCSn  25 MPa
r  27° r  33° r  33°

o  5.5 MPa o  7.4 MPa o  8.4 MPa



o  0.55
o  0.74
o  0.84
n  5.3 MPa n  7.0 MPa n  7.1 MPa

n  0.53
n  0.70
n  0.71

estimating gouge-production effects on reduced perme-


ability enhancement with shear and dilation. This will
result in less extreme increases in permeability with
shear unless the rock is hard, and/or the roughness low
and the rock hard enough to resist ‘wear’.
There may be an approximately 50% reduced con-
ducting aperture caused by gouge, according to Makurat’s
CSFT (coupled shear flow test) results. (Makurat et al.,
1990). This will be depend partly on the n /JCS ratio
which describes one of the important components of
damage. Damage during shear will also depend on JRC,
as steeper asperities tend to get more damaged, though
also cause more dilation. (Barton and Choubey, 1977
developed a damage coefficient on the basis of the above).
These opposed tendencies obviously have complex conse-
quences on the permeability development with shear.
As discussed before, both smaller scale and larger scale
joint or fracture non-planarity, will cause several possible
interpretations of ‘fracture orientation’, in relation to shear
wave splitting and polarization. Interpreted anisotropy
results that are actually non-parallel to H max but not
greatly different, might sometimes mislead analysts into
Figure 16.52 Experimentally determined reductions in JRC and thinking that the productive fractured reservoirs they are
JCS, based on analysis of the extensive scale-effect interpreting, have ‘open’ fractures approximately ‘paral-
investigations of Bandis 1980. Note the use, and lel’ to H max, when an alternative interpretation is pos-
important influence, of small-scale roughness JRCo sibly more realistic in geomechanics terms. (Here we
in this scaling. Bandis et al., 1981. recall from Chapter 15 that two conjugate sets can pro-
vide components suggesting H max parallelism.).
aperture increase, following Barton et al., 1985. The As demonstrated in the above BB-modelling exercises,
reality from case to case is difficult to estimate, without that parallelism of ‘open’ fractures to the H max direction
using the Olsson and Barton, 2001, JRC : JRCmob is unlikely in weaker rock unless joints are quite rough, or
modified, E to e conversion (Figure 16.9), for roughly production very poor, despite the fracturing (i.e. if the
536 Rock quality, seismic velocity, attenuation and anisotropy

joints were nearly closed they would satisfy the geome-


chanics criterion, but not the production requirement).
In stronger (e.g. carbonate) rock, with greater roughness,
there may be naturally larger apertures, which would give
‘open’ fractures also without the benefit of historic or
recent tectonic shear deformations.

16.8 Non-linear shear strength


and the critical shearing
crust
(a)
Hydraulic fracturing based stress measurements per-
formed for a pressure tunnel design in 1980 in Georgia,
USA while working in the USA, and acoustic emission
events recorded by others when injecting water into the
Cornwall Geothermal Project (Pine and Batchelor,
1984), were the stimulation for the sheared-fracture/
sheared-joint sketches given in Figures 16.53a and
16.53b, which we will use to introduce the ‘critical
shearing crust’ concept. These are ‘forced’ events, caused
by pore-pressure increase. Nature appears capable of
something similar without human intervention.
When there is a significant mismatch of horizontal
stress magnitudes, and pre-existing joints that are not
parallel to H, there is a good statistical chance for joint
shearing, when attempting hydraulic break-down in a
minifrac stress measurement (Barton, 1981). In a leak-
ing pressure tunnel-liner situation, the same event is
possible at much larger scale.
In this section, the important findings of Colleen (b)
Barton and Zoback and co-workers will be reviewed
Figure 16.53 a) The nature of conducting ‘open’ joints or
regarding the convincing evidence for the existence of
hydraulic fractures caused by pore pressure and shear
significant shear stress on joints or fractures that were stress induced shearing. Barton, 1986. b) The sug-
hydraulically conducting in a deep boreholes in crys- gested mechanism of hydraulic fracture ‘breakdown’
talline rock. Data from several wells are now available, by shear mobilization, given by Barton, 1981, for
thanks to the analyses of the Zoback team. By careful the case of pre-existing joints where stress measure-
analysis of fracture orientation and inflow logging, they ments (or leaking pressure tunnels) are involved.
also showed that insignificant shear stresses were acting
on joints that were not conducting.
In this work, they compared their stress estimates 16.8.1 Non-linear strength envelopes
with the ‘standard’ 0.6 
 1.0 joint friction model and scale effects
that has been popular in the USA since Byerlee, 1978.
(The ‘Byerlee law’ suggested a generally applicable In an attempt to understand the actual non-linearity of

 0.85 for the coefficient of friction). Before review- shear strength envelopes, both for rock joints, for frac-
ing their flow or no-flow data, we will look at the actual tured triaxial samples at high stress, and for intact rock
non-linearity of shear strength, in all its forms. Their itself, Barton, 1976 assembled and analysed a large
data on conducting joints and fractures fits the Byerlee number of test data from the literature. The following
‘band’ of shear strength, but the fit can be improved by contrasting sets of (non-linear) shear strength data,
considering non-linear friction. consist of high pressure triaxial shearing of (small)
Joint stiffness and compliance and the joint shearing mechanism 537

Returning to these and further data sets in 1990, at


the First International Workshop on Scale Effects in Rock
Masses in Loen, Norway, and armed with some import-
ant scale-effect research by Bandis, 1980, as sum-
marised in Bandis et al., 1981, the writer (Barton,
1990b) sketched approximate ‘scale effect’ bounds on
the significantly updated data sets.
Figure 16.55 shows the assumed ‘shifts’ in the data
boundaries. It was estimated that in case a) due to the
intensely high stresses to 800 MPa, there would be a very
limited scale effect (assuming faults were excluded). The
engineering data at stress levels mostly below 4 MPa
(a)
shown in Figure 16.55b, were known to display a severe
scale effect. In other words tests on 100 mm samples of a
given joint were always of higher strength than 1000 mm
samples, unless roughness was virtually absent.
Subsequently, the generally expected non-linearity of
the shear strength envelopes for rocks, rock joints and
crushed rock were assembled by Barton, 1999a, to com-
pare the various empirical strength criteria that have the
form:

 n tan(X log10 Y  .) (16.13)

where X represents the effect of roughness, e.g. JRC,


and Y represents the strength/stress ratio, i.e. joint wall
strength JCS/ n in the case of engineering stress levels,
which will need the confined strength ( 1  3)  JCS,
i.e. the strength/stress ratio ( 1  3)/ n , in the case
of shearing at high stress levels (Barton, 1976). The
schematic arrangement of the different strength
envelopes is shown in Figure 16.56.
Two of the sources for the schematic envelope for intact
(b)
rock in this figure, are reproduced in Figure 16.57. These
show the tendency of the numerous, mostly ultra high-
Figure 16.54 a) Shear strength of induced fractures in small, stress data to cross the ‘critical state’ line (defined by
extremely high pressure triaxial tests following failure Barton, 1976), in a more-or less horizontal orientation,
of the centimetre-size intact samples. (Data from
i.e. beyond the brittle-ductile transition, in the advancing
Byerlee, 1978 and others). b) Shear strength from
‘plastic’ state. Note that Z  c  26.6° in this model, as
direct shear tests on natural rock joints (one test per
sample), at low rock engineering stress levels. Typical
indicated by the expanding and then declining Mohr cir-
sample lengths of 100 mm to 1000 mm. Barton, 1976. cles in Figure 16.58.
The general non-linearity of the Barton-Choubey
(small-scale) and Barton-Bandis (general scale) constitu-
‘faulted’ samples, shown in Figure 16.54a, and a low tive equations for the shear strength of rock joints, can
pressure set of data shown in Figure 16.54b. The latter be reduced to the traditional, but actually less geotech-
were obtained from direct shear tests on natural rock nically representative Mohr Coulomb parameters c and
joints, and included the test results for the 130 fresh , by the simple equations shown in Figure 16.59. The
and slightly weathered joint samples tested by Barton particular case shows how to derive c and  between
and Choubey, 1977. arbitrary effective normal stress levels of n3 and n4.
538 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.55 a) Triaxial, high stress and b) direct shear (low stress) data for induced fractures and joints respectively. Scale effect curves are
estimates only. Barton, 1990b. Adapted from Barton, 1976 with additional data from Byerlee, 1978.

The general strengthening-by-confinement that we in weaker sedimentary rock, due to their limited toler-
have seen in the above shear strength envelopes, will obvi- ance of very high confinement, before ‘state change’,
ously apply strongly in the case of deep wells in hard crys- such as pore collapse (in chalks), or general porosity
talline rock, both for the intact rock, and for the faulted reduction. (Refer to Figures 12.22 and 12.23 for the
discontinuities or sheared joints. We will address this case of sandstones). These medium weak rocks would
aspect again when reviewing deep permeability data. obviously not support 20 to 50 MPa effective stresses,
Strengthening-by-confinement is likely to be of only and mostly maintain porosity, without some strengthen-
moderate strength in the case of hydrocarbon reservoirs ing due to confinement.

Figure 16.56 Schematic representation of the non-linearity of strength envelopes for various ‘earthscience’ materials, showing a comparison of the
empirically derived roughness (X) and strength/stress (Y) ratios. Barton, 1999a.
Joint stiffness and compliance and the joint shearing mechanism 539

Figure 16.58 A general model for representing the shear strength


of intact rock, showing the succession of key Mohr
circles that also span the brittle-ductile transition
and reach the ‘critical state line’, defined by Barton,
1976 as 1  3 3. From this point, strength envelopes
(a) may have a tendency to show reduced strength with
increasing confinement, in the ‘cap’ region.

24
2000

1800 23
12

1600
DIFFERENTIAL STRESS σ1σ3 MN/m2

1400 14
21

1200
20

1000 22 16 17

19

800 3σ 3

15 σ1
13 AL
IC
R IT TE
600 9 C STA

7
8
400
11
6 5
4
200 10

1
3
2
100 200 300 400 500

(b) CONFINING PRESSURE σ3 MN/m 2

Figure 16.59 Conversion of the non-linear Barton-Bandis shear


Figure 16.57 Shear strength data for intact rock at a variety of strength equation to Mohr Coulomb parameters c
mostly high confining pressures, from the numerous and  at a specific range of stress. Note use of large-
sources reviewed by Barton, 1976. scale rock joint characteristics, i.e. in situ block size
Ln with JRCn and JCSn, compared to Lo with JRCo
and JCSo for lab tests.
540 Rock quality, seismic velocity, attenuation and anisotropy

That there are limits to the strengthening-by-


confinement in the case of weaker rock, can be seen from
the ‘weakest’ strength envelopes in Figure 16.57, which are
seen in diagram b) to cross the ‘critical state line’ at com-
paratively low confinement. There might also be the occa-
sional ‘adverse’ effect of water flooding on e.g. the Ekofisk
chalk, where the confined strength was indeed exceeded in
the most porous sections, due to draw-down causing pore
collapse, but with excellent compaction drive production
as a result of the subsequent further weakening by water.
Before leaving the failure of intact rock it is of interest
to see the dilation and brittle fracturing that occurs when
rock is highly stressed. This was found by Rummel et al.,
1978, to be a significant source of P-wave anisotropy.
They utilised a biaxial loading arrangement with fast-
reacting servo-control, to study the development of dila-
tion adjacent to the shear failure surfaces. They found
that the P-wave velocity increased continuously in the
direction of maximum compression in the pre-peak
region. In the post-failure region, the P-wave velocity
decreased almost reversibly with reducing compression.
By comparison, the minimum and intermediate
principal stress directions suffered a marked reduction
of P-wave velocity (recorded as travel time increases),
after fracturing was initiated. (Enhanced permeability
would be a related phenomenon of such dilation: this
would presumably occur mostly in the 1 direction).
The authors mention the need to be aware of the possi-
bility for increasing velocity anisotropy, when interpret-
ing field seismic data in crustal regions where large
tectonic stresses are assumed to be operating.
Figure 16.60 shows the complete stress-strain curves
for their granite specimens, which were tested in a
1.0 GPa pressure vessel. These triaxial tests were on cylin-
drical samples. The markedly non-linear shear strength
envelope resulting from these highly confined samples
is shown in the same figure, together with the frictional
strength of the resulting fractures, which are also non-
linear, just as they were in Byerlee, 1978.
The frictional strength of the resulting fractures, fol-
lows the non-linear shear strength model of Barton,
1976 closely, with an equivalent JRC of 20, and the use
of confined strength ( 1 – 3) in place of JCS, as indi-
Figure 16.60 Shear stress – axial strain data for high pressure triaxial
cated in Figure 16.56. This was also the case when match-
testing of intact granite, showing strength envelopes
ing equivalent sets of data for various rock types, given
for subsequent sliding on the ‘minor faults’, and
by Byerlee, 1978. The friction coefficient (approx. 0.7) stick-slip using some artificial saw cut surfaces of the
for the Rummel et al., ‘fractures’ compares well with granite. Note the reducing non-linearity of the
the ‘critical shearing crust’ well data of Zoback and co- strength envelopes for these three categories. Rummel
workers, as we shall see shortly, although
 0.7 is et al., 1978.
Joint stiffness and compliance and the joint shearing mechanism 541

(a) (b)

Figure 16.61 Distinctive pole populations of joints or fractures that were conductive, and those that were not, from analysis of Cajon Pass
Scientific Drillhole results. Colleen Barton et al., 1995.

lower than some of their data. Note the lower friction near San Bernadino in California. The hydraulically
coefficients of the saw-cut surfaces, which resemble active fractures were located by temperature anomalies,
minor faults at residual strength. and mapped by televiewer, to an accuracy of 1.0 m
from each temperature anomaly.
The above authors had noted break-out direction
16.9 Critically stressed open anomalies over large depth intervals, in relation to the
fractures that indicate perpendicular-to- H max direction norm. These proved
conductivity to be associated with slip on joints (or minor faults).
Their plots of normalised mobilization of friction:
A significant contribution to our understanding of the ( / v) versus ( n – Pp)/ v shown in Figure 16.62 are a
geomechanics acting at deeper levels below the earth’s convincing demonstration of the importance of shear
surface, was made by Zoback and Colleen Barton and co- stress (rather than just minimum normal stress) for
workers, regarding the delineation of fracture directions explaining the ‘openness’ of joints, and therefore their
that appear to be conducting directions, and those that conductive capacity.
do not. The ‘simple’ yes or no concerning conducting One may speculate that in such granites and granodior-
directions is fundamental evidence, also for geophysi- ites, the values of JRC and JCS that one might typically
cists and petroleum engineers. estimate, would probably be high enough even for joints
Here we will review several cases that support the phil- that were under h min normal stress, to be marginally con-
osophy of ‘open’ joints actually often being those that are ductive. However, shearing-enhanced permeability would
under a significant state of shear stress, rather than being obviously greatly dominate in signal strength, as apertures
parallel or sub-parallel to H max and without shear stress. would be very small in the case of the ‘closed’ joints.
Authors in geophysics and even those working in the Such a proposition can be tested. A quick 1-D
petroleum industry, who are each dependent on ‘open’ Barton-Bandis modelling, starting with equation 16.11,
joints: both for shear wave anisotropy, and for petroleum with JRCn  5 or 10, JCSn  100 MPa, and an effec-
production, appear to have favoured the ‘parallel to H max’ tive normal stress of 75 MPa suggests conducting aper-
model. tures as small as 2.5 and 15
m. Apertures appear to
Figure 16.61 shows the differentiation of fracture reach such ‘residual’ levels after a few tens of MPa, due
orientations a) for those that were hydraulically conduc- to the combined effect (in this crystalline rock case), of
tive, and b) for those that were hydraulically non-con- high wall strength and significant joint wall roughness.
ductive. These different populations were recorded by Comprehensive supplemental data to the above deep
Colleen Barton et al., 1995 in the Cajon Pass Scientific borehole interpretation was more recently reported by
Drillhole, which was drilled to 3.5 km depth into gran- Townend and Zoback, 2000, and Zoback and Townend,
ites and granodiorites, 4 km from the San Andreas Fault 2001. These very extensive data sets are reproduced in
542 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.62 Mohr stress representation of the shear and normal stress components, acting on the Cajon Pass fractures, (i.e. sheared joints or minor
clay-free faults?), according to their conductive or non-conductive characteristics. Note authors’ friction coefficient limits, as per
Byerlee, 1978, (
 0.6 to 1.0). These could apparently be extended/modified to 0.4 to 0.9 in practice. Colleen Barton et al., 1995.

Figure 16.63 Normal and shear stresses for fractures identified as hydraulically conducting (closed symbols) or non-conducting (open sym-
bols). Cajon Pass (triangles), Long Valley (circles), Nevada Test Site (squares). Townend and Zoback, 2000, with data also from
Colleen Barton et al., 1995. (Zoback, 2006 pers. comm., by kind permission.) See Plates 14, and 15.

Figure 16.64 Normal and shear stresses for fractures identified as hydraulically conducting or non-conducting, using borehole imaging.
Cajon Pass (red diamonds and dots), Nevada Test Site (green circles and dots), Long Valley (yellow triangles and dots), KTB
(Germany – blue squares and dots). Inset shows
 / n for combined data set. Zoback and Townend, 2001, with data from
Ito and Zoback, 2000, and from Colleen Barton et al., 1995. (Zoback, 2006 pers. comm., by kind permission.)
Joint stiffness and compliance and the joint shearing mechanism 543

Figures 16.63 and 16.64 by kind permission. Note the


widely different stress magnitudes in these two figures.
Figure 16.63 shows conductive (closed symbols) and
non-conductive (open symbols or dots) from Cajon Pass
(triangles), Long Valley (circles) and Nevada Test site
(squares). It is clear from this data that either due to joint
character or due to stress level, or a combination of both,
the interpreted friction coefficients can range from (at
least)
 0.4–1.0 (Yucca Mountain: at normal stresses
below 16 MPa), to
 0.5–0.8 (Long Valley: 10 to
40 MPa), to
 0.5 to 0.9 (Cajon Pass: 14 to 70 MPa).
The quite wide range of
for Nevada Test Site (i.e.
Yucca Mountain) can be due both to the lower stress
levels (i.e. shallower depth), and to the wide range of
joint roughnesses. Joint roughnesses have been inter-
preted to be from about 1 to 15 for JRCo, when assum-
ing, for simplicity, a representative JCS of 100 MPa,
and r of 30°. These estimates are based on in situ
roughness recordings in the exploratory TBM tunnels,
and on a series of medium scale direct shear tests.
Values as low as
 0.4, shown in Figure 16.63,
implying a 22° angle of friction, suggest some clay-
fillings. In fact even lower values are indicated in the
Nevada Test Site data set, but these are non-conducting,
(a)
as expected for clay-filled discontinuities.
A clearer picture of the frictional content of the stress
data from these six deep boreholes, is reproduced in
Figure 16.65a and 16.65b, from Townend and Zoback,
2000, and Zoback and Townend, 2001. The crosses
shown in the figures are the error bars for the different
sets of data. The three dotted lines representing
 1.0,
0.6 and 0.2 were derived by these authors from the
Jaeger and Cook, 1979, equation:

S1  Pf
( )
2

2  1 
(16.14)
S3  Pf
(b)

The authors Townend and Zoback, 2000, and Figure 16.65 a) Differential stress versus effective mean stress and
Zoback and Townend 2001, do not give specific explan- b) maximum effective stress versus minimum effective
ations for the likely reasons for the above ranges of friction stress, for six sets of deep borehole data. See Townend
coefficient (notably 0.6 and 1.0). As interpreted, these and Zoback, 2000, and Zoback and Townend, 2001,
and other referred authors who are the sources for
are the levels of mobilized friction tolerated by the
the various data sets.
analysed joint planes, at the time of analysis. A man-
made change of pore pressure would not change the
differential stress, but would change the effective mean It is notable that the Cajon Pass rocks showed consist-
stress or the effective normal stress, thereby altering the ently high values of (resistance to)
 1. As shown
shear resistance of the joint planes in question, and earlier in this chapter, it is likely, in the case of joints
their friction coefficients, in view of non-linearity and fractures, that this is due to the interaction of a
with stress. moderate to high roughness JRCn, medium to high r
544 Rock quality, seismic velocity, attenuation and anisotropy

(say 30° approx.), and tolerable stress/strength ratios Table 16.16 Estimates of potential tolerance of mobilized frictional
n/JCSn. The high strength rock will give high values strength, to explore the meaning of
 1 for e.g.
of JCSn, or more specifically: high confined strengths Cajon Pass data from Zoback and Townend, 2001.
( 1 – 3), if joint walls had minimal alteration. Case A (minor alteration of Case B (hard unaltered joint
We can ‘assemble’ estimates of input data for applica- joint walls) walls)
tion in the high-stress, large block-size, version of the
JRCn  8, JRCn  10, ( 1 – 3)n 
Barton-Bandis criterion (Barton and Bandis, 1990):
JCSn  200 MPa, r  29° 400 MPa, r  31°
n  30 MPa
 0.72 n  30 MPa
 0.91
  ( 1  3 )   n  60 MPa
 0.65 n  60 MPa
 0.82

 n tan  JRCn log  n

  r
 (16.15)

  n  
Substitution of 500 MPa or 600 MPa for ( 1 – 3)n in
where ( 1 – 3)n is an estimate of the confined strength the Case B example in Table 16.16, again at effective
of the rock, with assumed absence of alteration along stresses of 30 and 60 MPa respectively, would raise the
the joint or fracture walls. If there was alteration and Case B estimates of peak
from 0.91 to 0.94 (with
reduced wall strength, one would expect a less ‘con- 600 MPa), or from 0.82 to 0.87 (with 500 MPa). Here
fined’ estimate of JCSn, i.e. of magnitude  ( 1 – 3)n we have ignored the inevitable small size of the samples
for the triaxially confined intact rock. Of course in situ, used to generate the Figure 16.57 data (down to approx.
the rock is poly-axially confined. 10 mm size cylinders), and have therefore assumed lim-
The following assumptions will be made to explore if ited scale effects at these intensely high pressures.
the toleration of transformed shear and normal stress in
a ratio as high as 1, i.e.
 1, can be ‘explained’ with 16.9.1 The JRC contribution at
a representative range of parameter values, as follows: different scales and
deformations
JRCn  8 or 10 r  29°–31°
JCSn  200 MPa or ( 1 – 3)n  400 MPa n  30–60 MPa
The only other conceivable contributor to higher shear
This is an effective stress range similar to the easily strength values might appear to be roughness JRCn.
estimated vertical effective stress at (hydrostatically However, with block sizes of 0.5, 1.0, or even 5.0 m,
saturated) depths of 2 and 4 km, assuming a simple the empirical scaling laws would demand quite excep-
average crystalline rock density of 2500 kg/m3 (i.e. tional JRCo values for the small-scale roughness, and
2  2.5 – 2 : 30 MPa, 4  2.5 – 4 : 60 MPa). As will frankly, little possibility that ‘the critical shearing crust’
be demonstrated, unless in situ block sizes around the was ever mobilized. The optimistic value of JRCn  10,
analysed wells are quite small, in fact significantly less applied in Case B in Table 16.16, is already actually ‘too
than 0.5 m, then impossibly high, lab-scale JRCo values high’, since it implies JRCo values (for nominal 100 mm
would be needed to generate mobilized friction as high scale), as high as 20 or 21, even when the assumed in
as
 1. One would need to greatly exceed the ‘JRCo situ block size round the analysed wells is limited to
limit’ (about 20), illustrated in Figure 16.26. 0.5 m. This is shown, by regression, from the inverse
As one may note from the estimates in Table 16.16, a application of the previously described block-size scal-
range of
 0.65 to 0.9 (approx.) can be reasonably ing of roughness:
‘explained’, but specifically from peak shear strength esti-
 L 0.02 JRCo
mation. If we engage in further analysis of maximum JRCo ≈ JRCn  n  (16.16)
possible confined strengths ( 1 – 3)n by looking at  L o 
more of the available triaxial strength data (Figure
16.57), we see the possibility of even higher strengths The standard set of ten JRCo values and their rough-
than 400 MPa at appropriate depths to our example, if ness profiles, from Barton and Choubey, 1977 were
the rock is very strong under triaxial conditions. reproduced in Figure 16.26. The ten selected samples
(Unfortunately for the ‘parallel to H max model’ weaker were mostly granite, aplite and gneiss, plus basalt, horn-
reservoir rocks do not have such benefits from confin- fels, slate and calcareous shale. The roughest sample no. 10
ing stress.). was an artificial tension fracture in a weak soapstone.
Joint stiffness and compliance and the joint shearing mechanism 545

The exceptional roughness needed to explain a toler-


ance of
as high as 0.94, namely: JRCo  20 if Ln is
0.5 m, therefore JRCn  10, together with the highest
defensible confined strength of 600 MPa (JCSn  ( 1 –
3)n), is still below the interpreted values of trans-
formed in situ principal stress anisotropy, of mostly
/ n 
 1. This was for the Cajon Pass research
borehole data of Zoback and Townend, who had inter-
preted effective mean stresses up to almost 60 MPa, at
3.5 km depth. The only possible explanation seems to
be that there was a small average block size, even at sev-
eral kilometres depth, in this tectonically disturbed area.
Two additional perspectives concerning roughness
are provided here, in view of the important potential Figure 16.66 Profiles and interpreted large-scale roughness coeffi-
contribution of large-scale roughness in explaining high cients (JRCn), from Bandis, 1980 scale-effect inves-
values of interpreted, transformed anisotropic stresses. tigations with model materials cast as joint replicas.
Details are also given by Bandis et al., 1981.
The even larger-scale roughness profiles shown together
with appropriate JRCn values, in Figure 16.27, were
obtained from extensive characterization of fractures 16.9.2 Does pre-peak or post-peak
developed in sandstone, tuff, concrete and hydrostone strength resist the assumed
samples, with sheared areas of about 1.0 m  1.10 m. crustal shear stress?
Characterization included roughness profiling at differ-
ent scales, Schmidt hammer testing, tilt testing of the A further factor to be considered when evaluating and
diagonally ‘jointed’ samples (weighing some 2.5 tons), attempting to geomechanically ‘explain’ the interpreted
and final biaxial shear testing in a 1 m3 loading frame at
-mobilized values close to 1.0, is that it is clearly not
TerraTek, in Salt Lake City, as reported by Bakhtar and logical that the conducting joints identified in these
Barton, 1984. As will be discussed at the end of this deep borehole studies by Zoback and his colleagues,
chapter, great difficulty was experienced in shearing should be exactly at peak strength. This point or even-
some of the rougher samples. This difficulty may have tual plateau on the shear strength–displacement curve
parallels to the Zoback and Townend deep borehole is in many ways a singularity, involving some few milli-
interpretations. metres of shear in the case of in situ block sizes, and per-
Bandis, 1980, performed a remarkable series of scale- haps a millimetre or less in typical lab samples. A
effect tests, using repeatedly cast replicas of natural joints, wide-ranging review of such displacement data, from
which were tested at different sampling lengths, that Barton, 1982 is reproduced in Figure 16.67.
could then be compared with each other since the The data is divided into three size categories (roughly
material strengths and compositions were identical. 5 to 30 cm  lab-scale, 0.3 to 3 m  in situ test scale,
He presented a set of eleven roughness profiles, with 3 m  exotic tests and up-scaled models). Even for
equivalent lengths of between 5 and 15 m, together small samples there is a wide range of ‘strains-to-peak’,
with back-calculated JRCn values. These are shown in when expressed as a displacement-to-peak divided by
Figure 16.66. The various profile lengths in this up-scaled sample length, and expressed as a percentage.
format, represent typical medium to hard rock strengths. A very interesting point arises when considering
Inspection of the major undulations on some of whether the shear-stressed joints involved in the ‘criti-
these long profiles, and on some of the 1 m profiles cal shearing crust’ are at their peak, pre-peak, or post-
shown in Figure 16.27, suggest that significant shearing peak shear strength. Reference to the JRCmob concept,
of joints involved in the ‘critical shearing crust’ scen- which was shown in Figure 16.10, and to the shear
ario, would likely be limited to JRCn values somewhat strength-displacement-dilation worked examples shown
below 10. Otherwise there would be a combination of in Figure 16.25, reveals that the ‘necessary’ permeability
too much volume increase, developing too much increase enhancement due to the beginning of dilation, actually
in normal stress and stiffness, thereby causing too high occurs when roughness starts to be mobilized, and may
shear strength for the mechanism to be viable. already be active after a few millimetres of shearing.
546 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.67 Extensive test data review of displacement-to-peak (peak), from a wide variety of sample/block sizes, from laboratory and
in situ direct shear tests. Barton, 1982.

As can be seen, the larger the modelled block size the undulating, too much dilation would prevent further
‘later’ this dilation and assumed permeability enhance- shear, due to a massive build-up of normal stress.
ment occurs. When there is faulting, or hydrothermal alteration,
In view of the need for combined shear stress and with clay-filled or clay-coated discontinuities present in
permeability, to qualify for registration in the mobi- the immediate vicinity, the relevance of the parameters
lized-friction-with-flow diagrams (Figures 16.61 to JRCn and JCSn in the Barton-Bandis joint constitutive
16.65), the joints need to be sheared, but they actually equation appear, at least at first sight, to be limited.
do not need to be sheared very far. If they are rough and The reality may be no remaining roughness of note
Joint stiffness and compliance and the joint shearing mechanism 547

(JRCn : 0), but with the possibility of a residual


(i) value, to account for the added shear resistance given
by any remaining undulation, with effective inclination
angle (i) in relation to the mean plane. This could be
simply accounted for by applying the Patton, 1966
equation, possibly with a cohesion intercept (c) added
for the case of (obviously cohesive) clay-fillings:

 n tan ( r  i )  c (16.17)

As was shown by a wide review of the shear strength of


clay-filled discontinuities, the residual friction angles (r),
and therefore overall frictional strength relevant to such
features ranges, from below 10° to about 25°, depend-
ing on mineralogy and degree of over-consolidation
(Barton, 1973b).
If one attempts to use the Barton-Choubey strength
criterion:  n tan [JRC log (JCS/ n)  r] to see if
it fits clay-related experimental data, it will be noticed
that when n (strictly n for effective stress) exceeds
JCS, (e.g. an over-consolidated fault gouge of 5 MPa
strength with a local effective normal stress of 20 MPa), Figure 16.68 Shear strength envelopes (peak, h  1%L, and
h  2%L) and assumed residual strength for direct
the log (JCS/ n ) term becomes negative. Contraction-
shear tests on model tension fractures of 100 mm
with-shearing, in place of dilation is then actually pre-
length. Barton and Hansteen, 1979. The material was
dicted by the model, (as for normally consolidated clay, a brittle, very fine-grained, sandstone-like material.
or heavily loaded over-consolidated clay), but not if The stress and strength scale:   400, and the length
JRC  0. Clearly it would be relevant to also apply a scale:  320, was due to lower model density of
suitably reduced value of r relevant to clay-smeared 2.0 gm/cm3. Note that
 0.6 and
 1.0 corre-
surfaces, to obtain a meaningful fit to data. Predictions of spond to the ‘residual’ strength envelope (0.58  tan
fault strength with this ‘rock joint and rock fracture’ 30°) and to the mean of the h  1 mm and 2 mm
constitutive equation are of course of limited reliability, strength envelopes, representing the shear strength
and there is little option besides testing or back-analysis. remaining (at different stress levels), following a mean
In further relation to the interpreted mobilized fric- 0.45 m of shearing at full scale, for these simulated
tion coefficients of 0.6 to 1.0 shown in Figures 16.61 to 30 m long fractures. However these samples were free
to dilate, with constant normal stress applied.
16.65, one may refer to direct shear test results on ten-
sion fractures developed in weak, brittle model mater-
ials, as reported by Barton and Hansteen, 1979. (These
materials could be formulated to give various frictional is then    m/r, where m and r are the respec-
and cohesive strengths, and consisted of fine sand, fine tive densities of model material and rock. With respective
Ballotini glass spheres, Pb304 red lead, and small pro- values of 2.0 gm/cm3 and 2.5 gm/cm3 (and   400),
portions of gypsum and water, with curing at high one can estimate a length scale  320. The samples
enough temperatures to cause disassociation of some of therefore represent approximately 30 m long, rough
the water: Ca.CO3.2H2O : Ca.CO3.1⁄2H2O thereby undulating major fractures.
reducing the compressive strength: Barton, 1971). Due to this scaling, the given shear displacement
It may be noted from the normal stress magnitudes magnitudes of 1 and 2 mm as recorded in the 100 mm
given in Figure 16.68 that there is a model/prototype long direct shear tests correspond to about 0.3 m and
(M/P) stress and strength scaling of 400, assuming a 0.6 m at full scale. The larger of these two shear dis-
strong 175 MPa prototype rock. The length scale from placements has reduced the shear strength of these
Buckingham’s -theorem for dimensionless products, rough, high strength simulated fractures (JRCo  20,
548 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.69 a) A dimensionless representation of the complete


shear stress-displacement behaviour of model tension
fractures shown in conventional format, in Figure
16.70. The shear test data has been normalised using
the JRCmobilized concept (Barton, 1980, 1982).

JCSo  0.44  400  176 MPa) to a minor fault char-


acter of
 1.0 at a simulated, medium normal stress
of 8 MPa, and to
 0.8 at a simulated maximum
Figure 16.70 Direct shear and dilation data for model tension
normal stress of 16 MPa, but each free to dilate.
fractures representing 30 m long, rough-undulating
The various data sets from Zoback and Townend,
fractures in hard rock. This is the source of DST
reproduced in Figures 16.64 and 16.65, are obviously data for the dimensionless representation in Figure
composed of variable rock strengths and variable joint/ 16.69. Barton, 1993.
fracture/minor fault roughnesses, since the different data
belong, obviously, to a wide variety of rock types, joint
characters, and minor fault types. These will inevitably It will be noted that the ‘ultimate’ strength, after
have suffered various deformation magnitudes, controlled some 1 m of simulated shearing (or 3% of the length of
by a variety of non-linear shear strength envelopes, repre- the fractures) is still well above the true residual behav-
senting different degrees of pre-peak, close-to-peak and iour, as indicated in Figure 16.69. This is also found
post-peak shearing, as suggested by the model tension when testing rock joints. It is not possible to reach true
fracture envelopes, shown in Figure 16.68. residual strength in the normal confines of a (linear)
In relation to post-peak shearing, it is helpful to invoke direct shear device. It is likely that Zoback and co-
the JRCmobilized concept again. (See Figure 16.10). Figure workers’ frictional interpretation of deep borehole ‘critical
16.69 shows a dimensionless set of shear strength dis- crust’ behaviour is showing
 0.4 to 1.0 coefficients
placement data, representing the whole range of behav- due not only to different JRCn/JCSn/r and n /JCSn
iour shown in conventional stress-displacement format ratios, but also due to different degrees of shearing,
in Figure 16.70. JRCmob can be estimated at any desired both pre- and post-peak.
location along shear stress-displacement curves, by evalu-
ating the mobilized shear strength ( mob) at the point of
interest. It is estimated by a simple re-arrangement of the 16.10 Rotation of joint attributes
Barton-Choubey peak shear strength equation: and unequal conjugate
jointing may explain azimuthal
  deviation of S-wave
tan1  mob   r
  polarization
n
JRCmob  (16.18)
JCS
log  We have previously investigated hypothetical joints
n
in typical reservoir ‘sandstone’ and reservoir ‘carbonate’
Joint stiffness and compliance and the joint shearing mechanism 549

(JCS  25 MPa and 50 MPa), and seen the dramatic data acquisition accuracy of each set of measurements.
closure of aperture that was predicted to occur if these There are many possible sources of error (Lynn, 2005,
joints were consolidated/closed by 10 or 20 MPa or pers. comm.), so this is uncertain.
higher effective normal stress, representing an assumed The logic and data of Zoback and Colleen Barton
range of (effective) h min, acting on sub-vertical joints and Townend and their colleagues regarding the prob-
or fractures in the reservoir. ability of shearing is surely incontrovertible. Although
Although down to a certain depth, an increased JCSn we perceived certain difficulties with
as high as 1.0 at
could be assumed, by utilising a confined compressive several kilometres depth, the ‘critical-shearing crust’
strength ( 1 – 3)n for the joint walls, it has been phenomenon was easy to verify on a local scale, using
demonstrated with triaxial test data that this increase is for instance non-linear shear strength and Barton-
limited in the case of weaker porous rocks. It may not Bandis coupled behaviour modelling.
be sufficient to explain ‘normal-closure’ apertures more Although presented by these authors in the context of
than a few microns, according to modelling predic- mostly multiple-kilometre deep boreholes in predomi-
tions. This is especially so if roughness is limited. nantly crystalline rock, it is the opinion of the writer that
With the advent of improved digital data acquisition, it is the petroleum industry that will have the greatest
and strong interest in the use of shear waves for inter- benefit of recognising the possibility (and sometimes
preting reservoir fracturing, it has become increasingly the necessity), of a shearing mechanism, often conjugate
important to correctly interpret what a certain level of in form. At Ekofisk in the North Sea, a model-predicted,
anisotropy and attenuation at a certain frequency means, and subsequently core-sampled confirmation of a con-
when analysing the fast and slow shear waves qS1 and jugate shearing mechanism (causing slickensided joints),
qS2. How far the industry has come in accepting the was probably fundamental for the two sets of 60° dip-
need for a shearing mechanism interpretation on some ping joints to be able to continue to drain the porous,
occasions, or a non-aligned set of fractures, or that two low permeability chalk matrix. (Nowadays there is an
intersecting fracture sets are actually the source of the additional component of compaction drive, due to water
measured components, is difficult to judge, even from weakening of the matrix and joints)
most recent publications. There would probably have been no development at
We should probably consider five possibilities: Ekofisk if there had been just one set of joints parallel to
H max. They would have been closed before discovery
● H max-aligned ‘open’ joints due to sufficient wall (i.e. therefore never discovered), and ‘further closed’ by
strength and roughness a 20 to 24 MPa increase in effective normal stress, (prior
● H max-aligned ‘open’ joints due to hard-mineral to water flooding), if an obstinate owner had decided for
‘bridging-but-not-blocking’ development. Of course other solutions (MHF and prop-
● ‘open’ joints at some angular deviation from the pants) would perhaps have been found effective.
H max direction So the details of fracturing, in terms of number of
● apparently ‘ H max-aligned-open-joint-set’ that actu- joint or fracture sets, may be important for production,
ally is two conjugate sets and therefore ideally should be detectable by seismic
● a generally non-planar joint wall topography in all interpretation. In addition, there is the simultaneous
cases importance of the non-planarity of most joints and frac-
tures. At one extreme, one has the cleavage joints in
The third and fourth of these cases will be addressed slates that give ‘billiard-table’ planarity and smoothness,
here. In the handful of cases reviewed in earlier chap- and clearly have no place in explaining petroleum pro-
ters, the anisotropic P-waves, or the polarized, split duction. They would ‘fail to produce’ on two counts:
qS-waves do indeed orient themselves in approximately their apertures would be too minute when acted on by
the ‘correct’ (conventional) direction in relation to the reservoir levels of effective normal stress, and in a ‘shear-
calliper-logged-perpendicular-to-break-out assumption ing earth’ scenario, where they would have difficulty tol-
for the H max orientation. However, we noted, e.g. erating
 0.4 when wet, they would nevertheless not
from Lynn et al., 1999, and Maultzsch et al., 2003 data, dilate when sheared. So permeability would be unlikely
that the line-up of assumed fracturing and assumed on both counts. At the other extreme there could be
major stress directions was not ‘perfect’. In fact there marked non-planarity of the one or two key joint sets, as
could be 10° to 20° discrepancies, if one can rely on the illustrated in the next figures.
550 Rock quality, seismic velocity, attenuation and anisotropy

Intermediate between these extremes of planarity or


non-planarity are all the other numerous cases, which
are vitally necessary for petroleum production because a
compromise between the extremes is needed for product-
ivity, if shearing is to be possible historically. There will
be no shearing (or only pre-peak shearing), if non-
planarity is too great, as suggested earlier when review-
ing the two sets of JRCn profiles.
Figure 16.71 shows several features that can explain
an interpreted ‘open fracture set’ orientation that differs
somewhat from the calliper assumed direction. The
undulating (non-planar) nature of conjugate joints that
have become slightly sheared, or even conjugate minor
faults, through successive shearing in response to H –
h (or S1 – S3) and perhaps fluctuating pore pressure,
has been exaggerated for clarity.
As has been noted previously in both Chapters 15 (a)
and 16, from the reconstructed shearing of model frac-
tures by Barton, 1973a, there is in Figure 16.71 an oppos-
ite ‘rotation’ of the fluid filled parts of the joints/minor
faults, shown with open O-symbol, in relation to the
rock-to-rock shear and normal stress transfer sections of
the asperities, which lie on the opposite ‘slopes’, and are
denoted by the R-symbol.
Open (O) and contacting (R) sectors, lie on either side
of the assumed mean plane of the joint (J). Note that
shearing of one of the conjugate sets has been assumed to
take ‘priority’, thereby displacing the secondary set, which
may also shear, but to a lesser degree. This is a frequent (b)
and almost inevitable consequence of ‘conjugate shear’. Figure 16.71 Idealised and exaggerated joint/fracture/minor fault
Considering the potential for shear wave splitting; if undulations, for a conjugate pair of ‘joints’, with one
we can make the assumption that the average orienta- dominant and displacing the secondary set. Note
tion of the fluid-filled sections (O) of the joints or minor contrary ‘rotations’ of ‘R’ and ‘O’ in relation to mean
faults have a different influence on the splitting/polariza- plane ‘J’. Shear wave splitting is assumed to occur
tion mechanism of shear waves, than the average orien- with different strength from the slip-prone ‘R’ sec-
tations of the rock-to-rock sections of the joints (R), tions, with their finite shear and normal stiffnesses
then we have immediately come 10° or more, further (and Kn  Ks), compared to that from the more
away from the H (or S1) direction. The ‘contrary’ fluid fluid-bearing ‘O’ direction. (Theoretical opinion is
rotations can be visualised from the ‘frozen’ evidence of sought here!) Reflection may occur more strongly
from the more fluid-bearing ‘O’ direction. Rotation
quartz injections into a shear stress field, causing actual
of the polarization of the qS1 and qS2 waves towards
shearing, as illustrated in Figure 16.72.
the H max direction would require the partial help of
As indicated in the extensive caption to Figure. an opposite component of polarization from the sec-
16.71, the influence of a second set of potentially shear- ondary set of joints, which might be less conducting
ing fractures/minor faults may be of importance in the and thereby have reduced lengths of fluid lenses. It
eventual polarization directions of the split qS-waves. If may be assumed that the pseudo-static shear stiffness
the dominant joint or fracture set physically displaces KS2 of the minor, displaced set would be increased in
the secondary set (as noted in the UDEC-BB model- relation to KS1 by the ‘off-set’ mechanism. This results
ling of Ekofisk joints, Barton et al., 1986), then we in an actual cohesive intercept in the sense of shear
have the possible influence of higher shear resistance resistance of the secondary set, further stimulating
and therefore higher shear stiffness for the secondary shear in the dominant set.
Joint stiffness and compliance and the joint shearing mechanism 551

Figure 16.72 a) Reconstructed shearing of a rough fracture at low


stress, showing the effective contrary rotation of
fluid-filled and rock-to-rock sections. b) A ‘frozen’
quartz filling of a minor fault showing the same con-
trary rotation of ‘fluid lenses’ and rock-to-rock con-
tacts. Høvik, Norway.

set. Presumably this could influence (reduce) shear com-


pliance in the secondary direction to some degree, even
though compliance is a dynamic micro-displacement
parameter.
Figure 16.73 shows a double-bladed guillotine method
for creating primary parallel fracture sets, and intersect-
ing sets of secondary parallel fractures in 25 mm thick
slabs of weak brittle model materials. These types of
interlocking ‘fractured’ models were used in pseudo-static
biaxial shear testing, as illustrated previously in Figure
16.42. They were also used to create ‘two-dimensional’
physical models of steep, ‘jointed’ rock slopes (Barton,
1971) with up to 40,000 blocks, and also to physically
model the excavation of large span, ‘near-surface’ cav-
erns, in various anisotropic stress fields, and with various
intersecting ‘jointing’ patterns (Barton and Hansteen,
1979). In each case the progressive deformation fields
were monitored by photogrammetry.
Figure 16.74 shows the results of direct shear tests
of three classes of model joint that have parallels in
nature. Peak and ultimate (larger deformation) strength Figure 16.73 a) Double-bladed guillotine for developing parallel
envelopes are shown for a) primary (set no. 1), b) pri- sets of tension fractures, and for creating intersecting
mary cross-cut set no. 1 (PCJ ), and c) secondary (stepped, sets by azimuthal rotation of the whole sample.
castellated) set No. 2. b) Continuous joints of set no. 1, and stepped/offset
joints of set no. 2, as often observed in nature.
Clearly the secondary set has a true cohesion inter-
c) Roughness profiles of the rough model tension
cept, and in a real situation at reservoir depth, would lie
fractures, as measured by photogrammetry. Barton,
somewhere between the shear strength envelopes for 1972a.
the intact material and the dotted ‘J’ curve in the Mohr
circle-based strength envelope diagram (Figure 16.58).
552 Rock quality, seismic velocity, attenuation and anisotropy

1968) that preceded the modelling of jointing in


UDEC and UDEC-BB, that started with the develop-
ments of Cundall, 1971.

16.11 Classic stress transformation


equations ignore the non-
coaxiality of stress and
displacement

In Figure 16.71a, there is discrepancy between the applied


shear stress orientation (S) parallel to the mean plane (J),
compared to the subsequent direction of shear displace-
ment when roughness is mobilized, and a non-co-axial
dilation begins. The resolved shear and effective normal
stresses ( and n) acting on the rock-to-rock contacting
parts (R) of the non-planar dominant joint, emphasise
this non-coaxiality. The non-planarity of most joints,
perhaps many minor faults, (and sections of major faults),
actually causes non-coaxial stress-and-displacement (i.e.
Figure 16.74 Direct shear test strength envelopes for primary, pri- violation of the coaxial stress and strain principle of St.
mary but cross-jointed and secondary joints, indi- Venant). This means that the ‘global’ application of the
cate that there will be higher shear stiffnesses, and Mohr Coulomb equation and following classic stress
therefore potentially lower compliance for the less transformation equations is actually erroneous to some
dominant directions. Barton, 1972a. varying degree.

The offset joints have shear strengths that are 1.5 to 3 1


 ( 1  3 ) sin 2 (16.19)
times higher than the primary joints, with greatest dif- 2
ference at lowest stress.
This offset-by-shear mechanism will obviously cause 1 1
the shear stiffness of this secondary set to be higher than n  ( 1  3 )  ( 1  3 ) cos 2 (16.20)
2 2
that of the primary set, and perhaps cause the dynamic
compliance of the secondary set to be lower than that of
the primary set. In the case of conjugate sets of vertical ( is the acute angle between 1 and the plane in ques-
reservoir jointing, this could be another potential tion. Note that equation 16.20 is sometimes quoted
source of azimuthal deviation of the shear-wave polariza- erroneously in literature, with a () sign between the
tion, away from the likely intersecting H max direction. two parts of this equation. A joint at 30° to the 1
The modelled (30 m long prototype) primary joints direction cannot have n  as implied in this particu-
gave measured ratios of Kn/Ks at prototype, scaled-up lar case: clearly  n).
stress levels as follows (Barton, 1972a). In theory the plane should be imaginary (thereby not
rotating the local principal stress directions), it should
n 3 7 11 MPa also not shear, and it should certainly not dilate, caus-
Kn/Ks 250 75 73 ing another error.
The problem with the classic stress transformation
At that time, in the late sixties, there was not the equations was experienced first hand by Bakhtar and
incentive to further investigate stiffnesses, apart from the Barton, 1984, who were attempting to biaxially shear
interesting problem that large anisotropy (or inequality) 1.0–1.1 m long diagonal fractures developed in 1 m3
of these two fundamentally different (pseudo-static) stiff- blocks in a large test frame at TerraTek, Salt Lake City, as
nesses reportedly caused numerical stability problems in referred to earlier in this chapter. (See large-scale rough-
the FEM joint element modelling (Goodman et al., ness profiles in Figure 16.27). In fact the same problem
Joint stiffness and compliance and the joint shearing mechanism 553

The discrepancy between the carefully estimated full


scale strength envelope, and the assumed loading path
(no. 1) is shown in Figure 16.75b. Correction for side-
friction and use of modified stress transformation equa-
tions, finally explained the ‘inexplicable’ high stresses
needed to achieve shearing. In fact most of the ten sam-
ples could not be significantly sheared, due to the clas-
sic neglect of the problem of dilation. The following
equations were needed to explain these experimental
difficulties:

(a) 1
 ( 1  3 ) sin 2 (  d n mob ) (16.21)
2

1 1
n  ( 1  3 )  ( 1  3 ) cos 2 (  d n mob )
2 2
(16.22)

1  JCS 
where d n mob  JRCmob log   n  (16.23)
2  n 

(The mobilized roughness concept of Barton, 1982 was


illustrated in Figure 16.10. The conventional equation for
peak dilation angle, uses JRCo or JRC n and not JRCmob.)
(b)
Example: Close to shear failure of the diagonally frac-
Figure 16.75 a) Large scale biaxial shear tests of 1 m3 blocks with tured block (Fig. 16.75):
diagonal, well characterized joints or fractures. Note 1) 1 had been increased to 28 MPa, 2 was reduced to
non co-axial ‘stress and strain’ due to dilation angle 0 MPa
dn. b) Calculated, scale-adjusted shear strength envel- 2)   45°
ope and 1) theoretical, 2) dilation corrected and
3) Assume dn mobilized  10°
3) fully corrected loading paths that explain the diffi-
culty of shearing due to a break-down of the classic With the ‘classic’ assumption of no dilation, equa-
stress transformation equations, when not modified tions 16.19 and 16.20 give:  n  14.0 MPa. (since
to account for the mobilized dilation angle. Bakhtar sin 90°  1.0, cos 90°  0). Shear failure of the frac-
and Barton, 1984; Barton, 1986; Barton, 1999a. ture had been expected/predicted even before ␮  1
was reached, as assumed here.
was experienced earlier, but not recognised, in the in With modified ‘dilating’ equations 16.21 and 16.22,
situ block test depicted in Figure 16.29. The limited the sin 110° and cos 110° terms (0.939, 0.342)
(0.25 mm) shear on this occasion was naturally indicate the actual difficulty in shearing, since 
assumed to be due to the ‘attached’ base of the block. 13.16 MPa (lower than assumed), and n  18.79 MPa
Ten 1 m3 samples with varying degrees of fracture (higher than assumed), dilation causing an effect simi-
roughness had been prepared and thoroughly charac- lar to an increasing angle .
terized, including the performance of large scale tilt- The new calculation of the actually applied ratio of
testing, following Barton and Choubey, 1977. The shear and normal stress (with an assumed mobilized dila-
experimental setup is schematised in Figure 16.75a. tion angle of 10°, was too low. With applied

Great care was taken in minimising side friction (S1 and only  13.16/18.79  0.70, the strength of the joint
S2 symbols in the figure), using double flatjacks and was too high for shearing more than a fraction of a
a thin Teflon sandwich separated by a fluid film of millimetre, i.e. pre-peak. Both these changes ( p, nq) in
molybdenum disulphide grease. the ‘actual’ compared to the ‘assumed’ boundary stresses
554 Rock quality, seismic velocity, attenuation and anisotropy

are also illustrated by the contrasting stress paths shown A long way beyond the peak singularity (where the
in Figure 16.75b. They help to explain the dangerous strength and dilation angle are maxima), there will be
burst of one of the real-life flat-jacks at 28 MPa pres- little need of a correction for dilation. So for mature
sure, nearly causing injury, and displacing pictures sections of faults one can apply the standard transfor-
around the laboratory walls due to the high pressure oil mation with almost no error, whereas in newer regions
burst. Only two of the surfaces illustrated in Figure of faulting, a roughness/dilation estimate may well be
16.27 could be sheared easily. required for more correct interpretation. Research in this
Since these experimental difficulties, further descrip- area, starting in the laboratory, is urgently needed.
tions and analysis of the problem have been given by
Barton, 1986, and Barton, 1999a, including convincing
evidence from the analysis of a large series of biaxial 16.12 Estimating shallow crustal
hydraulic-jack based shear box tests, where the applied permeability from a modified
loads S1 and S2 had to be resolved into the assumed com- rock quality Q-water
ponents and n acting on the 45°/45° oriented joint
planes in the centre of the shear box. (This is not Townend and Zoback, 2000, and Zoback and Townend,
required in a standard shear box, with perpendicular and 2001, argued convincingly for acceptance of a ‘critical
parallel application of forces in the principal planes). crust’ concept, with hydrostatic rather than lithostatic-
The main author of these biaxial shear box tests was related pore pressure throughout the brittle depth of
convinced that the JRC-JCS criterion of Barton and the crust. Their analyses of deep well data were
Choubey, 1977 needed to be modified. In fact the con- reviewed earlier in connection with the importance of
sistent ‘error’ in relation to a 1:1 gradient between ‘pre- the mobilized ratio of shear and normal effective stress,
dicted’ and ‘measured’ (as compared to an almost as a strong indicator of water-conducting fractures.
perfect 1:1 relation in standard shear boxes), was caused From the JRCmobilized concept we argued that perme-
by unintended neglect of a dilation adjustment in the ability enhancement could even be with pre-peak shear
classic transformation equations, as described above. displacements of just millimetre size. However the
On the basis of the above, Barton, 1999a argued that sometimes high ratios of resisted
( / n) of up to 1.0
‘classic’ stress transformation in all potentially dilating obviously suggests a close-to-peak condition, since we
geotechnical materials (e.g. rock joints, rock-fill, dense showed that it is difficult to construct such high in situ
sand, over-consolidated clay), may need to be con- full-scale shear strengths. Post-peak seems unlikely in the
sidered as violation of the St. Venant principle of case of the highest values of resisted
, while the values
co-axial stress and strain. below 0.6 to 0.7, implying mobilized friction angles less
One may speculate that the earlier referred Cajon than 31°–35°, could well be representing post-peak
Pass analysis of Zoback and Townend, indicating in this strengths, or minor faulting. Clearly even lower values
case, a tolerance of
 1, could also be subject to the of
of 0.4 to 0.5 (i.e. mobilized friction angles of only
possible need of a mobilized dilation correction, since 22° to 27°), as seen in Figures 16.62 to 16.65, have to be
this research borehole is located in the tectonically minor faulting, but without the ‘complication’ of sealing
active San Andreas fault area, with the possibility of a gouge, since they were also recorded as conducting
more tectonized ‘broken’ rock mass with smaller block- features, by the relevant authors.
sizes and perhaps rougher joints. As demonstrated earl- Zoback and Townend also assembled the deep bore-
ier in this chapter, it was extremely difficult to justify hole permeability measurements that are reproduced in
such high tolerance of shear stress, using all known Figure 16.76. In the case of the KTB hole in Germany,
means of maximizing shear strength. permeability test intervals varied between a few tens of
The dn mobilized equation 16.23, means that one must metres to 3.5 km, with bulk permeabilities between about
consider roughness (and therefore block-size), and the 5  1015 to 1017 m2, or approximately 5  108 to
magnitude of shear displacement. This is because the 1010 m/s if the water was close to 20°C.
magnitude of JRCmobilized depends numerically, upon In a shallower series of tests (1 km depth) at the
the degree of shear displacement. So the mobilized dila- Monticello reservoir, Zoback and Hickman, 1982, (with
tion angle will depend on the present displacement, in data also shown in Figure 16.76), reported permeabilities
relation to the initial, usually interlocked condition of ranging from 1015 to 1016 m2, or 108 to 109 m/s.
the joint. Thermal models of borehole temperature suggested
Joint stiffness and compliance and the joint shearing mechanism 555

As we have seen earlier, nearer the surface, the degree


of potential connectivity in a rock mass may be reflected
in the relative block-size ratio of RQD/Jn, because of the
great importance of the number of joint sets Jn.
However when clay was responsible for low RQD val-
ues, this relation would likely be compromised. One
began to speculate whether the second pair of shear-
strength related parameters, Jr/Ja could also have a role
in ‘explaining’ relative magnitudes of permeability for
individual joint sets?
We will start by describing the actual limitation of
what was proposed in Barton, 2002a, where an
extremely simple approximation between rock quality
and the Lugeon value was proposed for central ‘jointed
rock’ qualities in the integrated Q-diagram, reproduced
here for easier reference, as Figure 16.77.

L ≈ 1 Qc (16.24)

where L  Lugeon (1 Lugeon  107 m/s  1014 m2


at 20°C), and where Q c  Q  c/100, where c 
uniaxial compression strength of the rock.
As preliminary examples we see K  1014 m2
(107 m/s) when Q c  1 (a quite heavily jointed rock
mass, e.g. 3 to 4 joints sets, near-surface), and we see
Figure 16.76 Zoback and Townend, 2001, assembly of deep crustal
permeability from well-known deep borehole proj-
K  1016 m2 (109 m/s) when Q c  100 (sparsely
ects such as Cajon Pass, KTB, Kola. Intact samples jointed, quite massive rock, typical at hundreds of metres
(with K 1018 m2) are given on the left. See origi- depth and beyond). So far the approximation is reason-
nal paper for references to other projects. able, but as hinted earlier in this book, the relation
L  1/Qc inevitably breaks down when there is clay-
filling along the joints or fault zones.
1014 (107 m/s) for the 0 to 2 km depth interval at As may be noted from Figures 16.76 and 16.77, the
Kola, and 1017 m2 (1010 m/s) for the 6 to 8 km depth above deep well data (often 109 to 1010 m/s, or
interval. The above, and similar data, led Zoback and 0.01–0.001 L), lies in a partly ‘consistent’ position on the
Townend, 2001, to conclude that the upper crust had Q (or Q c) – Vp – L chart concerning the high quality end
permeabilities of 1017 m2 (1010 m/s) to 1016 m2 of the diagram, showing ‘hard jointed’ and ‘hard massive’
(109 m/s) over 1 km and 10 km (depth) scales. black-and-white curves. However, extrapolation of the
diagram to much greater depth would be needed, by
extending the ‘1000 m’ depth diagonal upwards, into
16.12.1 The problem of clay-sealed higher velocity and higher modulus territory.
discontinuities Recent compilations of permeability data from the
Äspö site in Sweden, which is a quite well-jointed site
With this interesting evidence for the maintenance of in diorite and granites, shows a range of K  104 to
some significant permeability to great depth, presum- 1010 m/s based on analysis of numerous tunnel probe
ably related with even deeper operation of the ‘critical holes from 50 to 400 m depth. There was also a recorded
shearing crust’ mechanism, it was of interest to know if 260:1 permeability anisotropy around the spiral access
there was a possibility that a rock mass description tunnel (Vidstrand, 2003). An interesting scale effects
scheme like the Q-system, could ‘explain’ such diverse analysis by the same author showed K varying from
ranges of permeability, and the partial maintenance to about 105 m/s to 108 m/s at the 100 m scale, and K
greater depth. varying from about 104 m/s to 1010 m/s at the 10 m
556 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.77 The jointed rock mass quality Q-diagram, showing potential integration of the UCS-normalised Q c value, Vp, Emass and
Lugeon value. (1 Lugeon  107 m/s). (See Chapter 9).

scale, Below the 1 m scale, values reduced to intact Table 16.17 Some suggested inter-relations between rock mass
matrix values of 1012 m/s as a minimum. quality, P-wave seismic velocity, and permeability. Note
The above six orders of magnitude corresponds to that ‘faulted rock’ qualities, with Q  0.1, implying
clay-filled (and probably sealed) discontinuities, have
the full range of permeabilities shown in Figure 16.77.
been excluded.
The simple (too simple) correlations between rock
quality, velocity and permeability, presently represented Extremely/
in Figure 16.77, are reproduced in Table 16.17. very poor Poor Good Very good
In Figure 16.78 two sets of equations are shown in Qc 0.1 1 10 100
the table above the figure. The uppermost set of equa- Lugeon 10 1 0.1 0.01
tions correspond to the L  1/Q c inverse relation dis- K (m/s) 106 107 108 109 (approx.)
cussed above, and show the standard ‘definition’ of rock Vp 2.5 3.5 4.5 5.5 (km/s)
mass quality Q, here normalised by the UCS of the
rock. The lower part of the table shows recent work by
that the friction-based Jr/Ja ratio has been inverted to
the writer for rationalizing the permeability – Q-value
Ja/Jr to allow improved consistency where clay-filling is
linkage, which has inconsistencies where clay-filled joints
concerned. With clay-filling on dominant jointing,
or discontinuities are concerned. One may note from
Qwater now increases, and the inverse permeability
the following:
reduces. A further step to enable general stress or depth
dependent permeability to be modelled, is normaliza-
RQD J J tion by 100/JCS; i.e. when the joint wall compression
Q water   a  w (16.25)
Jn Jr SRF strength is less than 100 MPa, there will be an increase
Joint stiffness and compliance and the joint shearing mechanism 557

Figure 16.78 The permeability-depth trends that are predicted by the QH2O modification of the rock quality Q formula. The trends are pre-
liminary, but show the potential for a comprehensive rock mass description method to give a first estimate of possible permea-
bility values.

in ‘quality’, and the inverse, giving permeability, will Example 1.


reduce further: Clay-bearing, well-jointed rock at 100 m depth, with a
low UCS of 10 MPa.
RQD J J 100
Q H2 O   a  w   
Jn Jr SRF JCS
(16.26)  regular Q-value  50  1.5  0.66 
 9 4 1 
  1.4, i.e. ‘poor’ 
Finally, there is a preliminary depth-permeability- 50 4 0.66 100
QH2O relation, based on analysis of shallow-borehole Q H2 O      98
9 1.5 1 10
Lugeon testing at a rail tunnel project in the Oslo area,  2 
and at a metro tunnel project in São Paulo. K ≈    9  109 m/s
 1000  98  100 5 3 

2 (Quite low permeability due to clay coatings, and com-


K ≈ 5 (16.27)
1000  Q H2 O  D 3 pressible joint walls).
Example 2.
The inclined lines in Figure 16.78 show this relation Sparsely jointed, rough undulating joints, quite massive
in graphic form, with an extended permeability range rock, with high UCS  200 MPa, with high water pres-
101 to 1011 m/s. It is relatively easy (perhaps too sure at 1000 m depth.
easy), with the above formulations, to ‘construct’ realis-
 
tic permeability estimates at depth, with any desired  regular Q-value  100  3  0.5 
jointed or faulted rock mass quality/character (see  3 1 0.5 
  100, i.e. ‘very good’ 
Appendix A for details of the Q-parameter ratings). So
100 1 0.5 100
far, the method is appropriate to ‘jointed and faulted’ Q H2 O      5.5
rock problems, and application in highly porous rock 3 3 0.5 200
 2 
types, with significant matrix permeability must be K ≈  5   4  109 m/s
avoided.  1000  5.5  1000 3 
558 Rock quality, seismic velocity, attenuation and anisotropy

Figure 16.79 a) An example of a massive rock mass with rock quality Q  1000 and a deformation modulus in excess of 100 GPa. A very low
attenuation is implied. b) The fault-collapse blocking the tunnel on the right would give almost the lowest rock quality
Q  0.001, and a modulus of deformation lower than 1 GPa. It is perhaps ‘off-the-scale’ regarding the conventional definition of
Qseis, and would need to be under stress to allow spectral analysis of measurable amplitudes. Its Q-value would then be higher too.
See Plate 16.

(Quite low permeability due to limited jointing and a particularly difficult tunnelling project. The different
high stress level, which cancels out the conducting effect qualities are typified rather well by the need for a cable
of hard, rough joint walls. Near-surface, significant per- car to reach the top of ‘Q  1000’ rock, while a boat
meability). was needed for hazardous tunnel inspections of several
similar ‘Q  0.001’ fault zones, due to the flooded
Example 3.
state of the tunnel between each fault blockage.
Negligibly jointed, low permeability, massive rock at
As we have seen in several chapters the seismic Q val-
5 km depth (here we extrapolate beyond the curves of
ues, the inverse of attenuation, might have similar num-
Figure 16.78, and base the estimate on the equations).
bers, remarkably, as these pseudo-static moduli (when
  expressed in GPa) namely 100 and 1.
 regular Q-value  100  4  1  While permeability might be even lower than
 1 1 0.5 
  800, i.e ‘extremely good quality’  0.001 Lugeon in the Sugar Loaf (i.e. 1010 m/s,
or 1017 m2), roughly corresponding to the Figure
100 1 1 100
Q H2 O      25 16.77 prediction, the fault zone quality of 0.001 would
1 4 0.5 200 be very unlikely to give 104 m/s or 1011 m2 close to
 2 
K ≈  5   5.5  1011 m/s the surface, as implied in Figure 16.77, because a lot of
 1000  25  5000 3  sealing clay is obviously present. If devoid of clay, Q
could remain quite low, and permeability quite high.
(extremely low permeability due to lack of jointing and Application of ‘QH2O’ from Figure 16.78 to this
high stress level) faulted case, with an assumed 0.1 MPa strength-of-clay
By way of illustrating rock mass Q-values and their JCS estimate near-surface, and 1 MPa at 1000 m, would
extreme range, and also rounding off this wide review give a predicted range of permeability of about
of ‘the rock beneath our feet’, a pair of contrasting rock 106 m/s at the surface – i.e. wet conditions, while at
conditions are illustrated in Figure 16.79. By chance 1000 m depth, a value closer to 109 m/s would be sug-
both are from Brazil, one a famous landmark, the other gested. These estimates appear reasonable.
17 Conclusions

Introduction PART I

In a great range of applications stretching from tectono- Chapter 1 Shallow seismic refraction
physicists’ interest in microcracks, frequency depend- and importance of
ent attenuation and earthquake source mechanisms, civil rock type
engineers’ concerns with low modulus dam foundations,
petroleum engineers’ interest in shear-wave anisotropy 1.1
and the permeability of fractured reservoirs, or tunnelling
engineers’ concerns with the approaching difficulties of a Many geophysicists insist that obtaining high-resolution
low velocity regional fault zone, the common use of seis- images from ground level to just 50 m depth is still one
mic measurements unites many fields of earth science. of the major challenges of modern geophysics. This
The richly illustrated material in this book has been happens to be the layer of the subsurface closest to most
assembled as a result of an interest in a variety of civil, of our civil engineering endeavours, from tunnels, to
mining, petroleum, geophysics and earth science fields. dams, to the foundations for high buildings.
The common denominator has been rock mass and rock
joint behaviour and their impact on the seismic inter-
pretation of the sub-surface.
1.2
The geophysics of the sub-surface and the rock
mechanics of jointed media often focus on related aspects
Shallow refraction seismic measurements – using first
of the same infinitely variable material. Yet different scales
arrival, compressional P-wave velocities close to the sur-
and different frequencies have caused these disciplines to
face can give a remarkable picture of near surface con-
have a mostly separate development, with limited cross-
ditions due to some fortuitous interactions of physical
referencing in the multitude of journals. Bridging this
phenomena. Weathering and the usual lack of significant
void in some strategic locations is the major objective of
stress near the surface has allowed joint systems, shear
this book. Two of the strongest bridges will be seismic
zones and faults to be exaggerated in both their extent
quality and rock quality, and fracture compliance and stiff-
and severity. Stress levels are low enough to allow joints
ness, as applying in particular to fractured reservoirs.
and discontinuities to be seismically visible due to their
Part I of this book is mostly focussed on civil engi-
measurable apertures. Acoustic closure occurs at greater
neering, and the links between seismic interpretation of
depths than those usually penetrated by conventional
rock conditions at laboratory and field-scale, and their
hammer seismic, unless rock strengths are rather low.
impact on rock quality interpretations for tunnels, deep
foundations, dams, planned nuclear waste repositories,
and mines.
Part II of this book focuses on greater depths, greater 1.3
scales, and more subtle geophysical detail, as befits this
rapidly developing field. The chapters of Part II treat atten- Micro-fractures and rock joints are sensitive to stress levels.
uation and anisotropy in detail, and topics range from the The more closed state of the discontinuities that are per-
use of shear wave splitting to interpret anisotropic frac- pendicular to the major stress, and the more open state of
tured hydrocarbon reservoirs, to the interpretation of those that are parallel will give the rock mass anisotropic
mid-ocean spreading-ridges, and crustal conditions as stiffness. The rock mass will therefore frequently dis-
interpreted between earthquake source zones and the play anisotropic seismic velocities. Hydraulic conduc-
near-surface. tivities and deformation moduli that show anisotropic
560 Conclusions

distributions should therefore be easily detected by scale of quality from about 0.001 (for major, clay-filled
near-surface seismic refraction measurements. fault zones) through 1.0 (for well-jointed rock) to 1000
(for massive, unjointed bodies of rock). Static rock mass
deformation moduli might vary from 0.1 GPa, through
1.4 10 GPa to 100 GPa for the same range of qualities, more
exact values depending on the strength and porosity of
The intimate interaction of the dynamic Poisson’s ratio the rock material, and on the level of stress or depth.
and the compressional P-waves and transverse S-waves, Qrock also has relationship to the seismic-frequency
gives strong correlation with the physical condition of based magnitude of Vp, with 0.5, 3.5 and 6.5 km/s cor-
sediments, due to the propagation of P-waves in fluids responding, roughly, to the above rock mass qualities,
and the lack of propagation of S-waves due to the absence when measured near-surface.
of shear strength in fluids. Soft, saturated sediments that
display Poisson’s ratios of almost 0.5 have negligible S-
wave velocities and large ratios of Vp/Vs as a result. This 1.8
ratio eventually falls to below 2.0 as Poisson’s ratio falls
and lithification occurs, as seen at greater depth. Shear The Q symbol is also used by geophysicists as a mea-
zones or faults display high values of Poisson’s ratio and sure of the attenuation or energy loss when seismic
higher ratios of Vp/Vs for similar reasons, despite the rel- waves propagate through a medium. Low energy stor-
atively low value of Vp in such zones. age and high energy loss per cycle, signifying low Qseismic
values and high attenuation (equal to the inverse
Q1seis ), will obviously correspond to low rock quality
1.5 Q-values, and vice versa, suggesting potential correla-
tion between these two Q symbols. As will be shown in
Shear-waves offer better vertical resolution than com- Part II, correlation may be most valid through a ‘shared
pressional-waves in shallow, unconsolidated sediments. earth’ parameter: the stress-dependent rock mass defor-
Shear-wave velocities in such cases are only a fraction of mation modulus.
the compressional-wave velocities. This results in very
small wavelengths, despite the fact that the dominant
frequency of shear wave data is generally lower than for 1.9
compressional waves. To obtain the same level of reso-
lution with P-waves, energy of very high dominant fre- The seismic refraction method has some important limi-
quency has to be generated, and this is correspondingly tations. Horizontal or sub-horizontal ray paths record only
more attenuated in the low seismic Q sub-surface. the upper part of each seismic layer. A thin high-speed
layer can mask underlying material, while a low-velocity
intermediate layer will not be recognised for similar rea-
1.6 sons. Depth calculations to underlying refractors will then
be erroneous. Hidden low velocity zones can be detected
Shear-waves are not attenuated at the water table, and by up-hole shooting from a borehole to the seismic
are little affected by changes in fluid saturation. They spread (i.e. reversed VSP), and of course by inspection
may therefore more easily detect lithological changes with and index testing of core, if boreholes are available.
correspondingly less ambiguous velocity contrasts. Under
conditions of full water saturation, P-wave velocity con-
trasts between similar lithologies may be small, whereas 1.10
shear-wave velocities may reflect true lithological changes.
The overburden at typical shallow refraction sites for
civil engineering projects may not present uniform con-
1.7 ditions. Even a simple site may contain velocity anom-
alies, which reduce the image quality. The time horizons
Rock mass quality can be described by the Q-value suffer push-down beneath slow-velocity anomalies, and
(here frequently termed Qrock), which is a logarithmic pull-up beneath fast velocity anomalies.
Conclusions 561

Chapter 2 Environmental effects on 2.4


velocity
A large majority of matrix velocity-strength data lies
2.1 between the trend lines c  V p3 and c  0.25 3p , with
Vp expressed in km/s. However, this 4:1 range is still
Numerous factors influence seismic velocity. Joint fre- insufficient to encompass some high porosity, low
quency, porosity, rock (and joint wall) strength, density, strength data, and is also insufficient to encompass some
depth, stress, stress anisotropy, degree of saturation, and high strength igneous rock data. c  2 V p3 may be
type of saturating fluid are among the primary influ- needed as an upper limit. A doubling of rock strength for
ences. Velocity generally increases with depth and with each 1 km/s increase in Vp is also a good mean trend at
rock stress increase, due to increased joint normal stiff- intermediate velocities. The effectively concave velocity-
ness, joint aperture reduction, joint frequency reduction, strength trend implied by the above, is replaced by a lin-
and reduction of clay content in the joints. The increased ear Vp – Schmidt (N) hammer rebound trend, with Vp 
depth also causes deformation modulus to increase, and 0.1 N as a first approximation.
in general, the permeability reduces due to the increase in
effective stress. There may also be changes of lithology
with increased depth. Seismic interpretation must there- 2.5
fore be neither separated from the geology, nor from the
structural geology. At the low end of the Vp- c spectrum ( c  1 MPa to
10 MPa), laboratory data for intact samples of Tertiary
mudstones and sandstones show roughly an order of
2.2 magnitude increase in strength (1 MPa to 10 MPa) for
1 km/s increase in velocity (i.e., 1.5 km/s to 2.5 km/s).
Surface weathering effects alter numerous properties of Concerning the rock mass, another general trend is
both the rock material and the rock mass. Reduced den- that rock mass quality Q-values increase 10-fold for each
sity, reduced compression strength and increased porosity 1 km/s increase in Vp for the case of hard, low porosity
correlate both separately and collectively with strongly rocks at shallow depth.
reduced matrix P-wave velocity. Because of the increased
void space in weathered materials, there is a marked
reduction in velocity as the degree of water saturation 2.6
reduces below 100%. There is a less pronounced reduc-
tion in velocity with reduced saturation, when rocks In the case of shale, with its mixed content of similar
are unweathered. Dry rock shows the greatest range of density minerals (quartz   2.66 gm/cm3, illite and
P-wave velocities. montmorillonite   2.61 gm/cm3 ), the correlation of
compressive strength and density is inevitably poorer
than for most rocks. P-wave velocity is therefore not a
2.3 sensitive indicator of compressive strength for shales.

Linear trends for Vp versus the inverse of porosity


(1/n%), and linear Vp versus density () relationships, 2.7
may conceal a non-linear Vp–uniaxial compressive
strength ( c) trend for the matrix. Weathering reduces When slopes or foundations are excavated, there is usu-
all three parameters, and therefore also the velocity. The ally a change of stress (mostly unloading) and a reduc-
approximately inverse proportionality between velocity tion in pore pressure. Improved drainage and increased
and porosity is influenced by many subtle variations permeability are also likely. Monitoring of Vp in such
bought about by such factors as clay-content in sand- zones usually shows significant reductions in velocity,
stones. Where joint or crack porosity increases due to especially when the rock mass is significantly jointed or
weathering, and the matrix porosity increases too, the damaged by the excavation process. Depending on the
reduction of velocity may be several km/s, causing very type of excavation, there may also be shear stress devel-
steep velocity-depth gradients near the surface. opment due to the excavation, which may accentuate
562 Conclusions

such effects. Weathering, and frost damage near the may be mainly a result of coring damage caused by the
surface, may accentuate these velocity reductions in sus- release of anisotropic stresses. Stress dependent behav-
ceptible materials, if left unprotected or insufficiently iour is particularly pronounced at low stresses compared
supported. Velocity will then tend to decrease over to the higher virgin stress. Above the previous stress
time, signifying reduced resistance to shear stress. state, the sensitivity to stress change is less. There is lit-
tle or no stress dependence when no cracks are formed
in the recovery process, or when the rock is loaded or
Chapter 3 Effects of anisotropy on Vp unloaded near the original stress state. In a limited
stress regime around the original stress state, the rock
3.1 matrix may behave as a linear elastic material.

Even in the unstressed state, and even if stresses are


isotropic, the presence of microcracks, fabric, bedding 3.5
or jointing will give anisotropic distributions of seismic
velocity if these features are themselves anisotropically Vertical or sub-vertical joint sets showing dominant strike
distributed, as is obviously the case for fabric and sedi- and continuity of the primary set, cause anisotropic P-
mentary structures. wave velocities. The degree of velocity anisotropy may
range from 10% for high velocity rock masses to as much
as 40% for lower velocity rock masses. The magnitude of
3.2 anisotropy is accentuated if the dominant set is also paral-
lel or sub-parallel to an anisotropic distribution of the
Intact specimens of rock that exhibit strongly anisotropic principal horizontal stress magnitudes H max and the per-
or orthotropic tendencies such as slate, show significant pendicular h min. Near-surface jointed chalk has shown
velocity differences when measured parallel to foliation velocity contrast in perpendicular directions as extreme as
(e.g., 5 km/s) and perpendicular to foliation (e.g., 4 km/s). Vmax. of 2.85 km/s and Vmin. of 1.75 km/s, giving a total
This anisotropy varies smoothly as the angle of incidence velocity anisotropy (Vmax.-Vmin.)/Vmax.  0.38 or 38%.
to the foliation is varied from 0° to 90°. Intact samples of
quartzite, amphibolite, hornfels, gneiss, phyllite, schist,
and slate exhibit ‘weak’, ‘moderate’ and ‘strong’ degrees 3.6
of foliation. They can be expected to show matrix
velocity anisotropies of 2–6%, 6–20% and 20–40% The commonly occurring interbedding of sedimentary
respectively. strata, such as sandstone, shale and mudstone, represent
layers that also have different porosity, density, modu-
lus and uniaxial strength. These contrasts of the major
3.3 mechanical parameters cause variable velocity anisotropy
for perpendicular and parallel wave propagation. The fine
Micro-cracks that are closed by stress give higher velocity layering of sedimentary strata means that the dominant
in the direction of the applied stress. Fabric such as schis- wavelength of a seismic pulse is long compared to the
tocity will also give velocities that are strongly dependent thickness of individual layers. The medium will exhibit a
on stress levels, when the loading is normal to fabric and vertical symmetry axis in the case of horizontal layering.
the rock is dry. The same reasoning applies to joint sets (For detailed treatment of anisotropy, see Part II).
that are closed by the stress or depth effect. When however
loading is parallel to predominant microcracks, fabric or
jointing, velocities are higher and there is less sensitivity to Chapter 4 Cross-hole velocity and
stress level, either in the dry or saturated states. cross-hole velocity
tomography

3.4 4.1

It has been suggested that stress dependent velocity caused Traditional cross-hole or between-adit velocity mea-
by microcracks seen in cores taken from great depth, surements at dam sites gives only average velocities for
Conclusions 563

use in extrapolation of deformation measurements. Cross- mine pillars, or through rock masses under high in situ
hole seismic tomography allows both the width and stresses, a clear effect of rock stress on seismic velocity is
location of low velocity zones to be estimated, and has demonstrated. In hard jointed rocks near the surface,
revolutionized sub-surface seismic measurements for the P-wave velocity may increase by 1 or even 2 km/s as
tunnels, caverns, deep foundations (and petroleum a result of a stress increase of only 5 MPa, and this may
reservoir monitoring: see Part II). occur where there is no rock mass quality improve-
ment. In soft jointed rock, acoustic closure may mask
the subsequent effects of a stress increase, unless higher
4.2
stress causes compaction of the pore space. Care needs to
be taken in the interpretation of higher velocities meas-
Severe topographic changes and gradational weathering in
ured at depth. These may or may not be associated with
mountainous terrain, make the use of conventional travel-
improved rock quality at depth.
time refraction seismic hard to use, as long geophone
arrays may receive shortest path direct waves earlier than
the refracted head waves. Gradational, progressive weather- 4.6
ing, rather than distinct layering, causes less clear develop-
ment of head waves. The use of tomographic inversion When the cross-hole tomography method is used to image
techniques for tunnels through steep terrain, is therefore highly stressed and burst prone areas in mining, steep
an attractive alternative, and may include tunnel to surface velocity gradients may be found associated with such
imaging, where fault zones are expected. zones. Attenuation tomograms that change with time as
mining advances, may be due to high stress anomalies,
4.3 stress release phenomena, changes of joint aperture and
stress induced fracturing. High shear stresses may be pres-
There is an apparent decrease in the velocity of high veloc- ent where steep velocity gradients occur. Seismic velocity
ity layers with increasing separation of boreholes. The high tomography can also be used to follow the effects of loos-
frequency direct first arrival received at small borehole sep- ening and void formation caused by blasting.
arations may be replaced by a long dispersed wave-form
at the largest separations. Attenuation of the higher fre- 4.7
quency, higher velocity part of the wave at increasing dis-
tance may occur in strongly attenuating rock masses. P-wave amplitude and S-wave frequency measurements in
the laboratory have shown superior sensitivity to stress
4.4 change and to the effects of joint frequency change,
as compared to P-wave velocity. Amplitude attenuation
When comparing cross-hole and downhole velocity meas- tomography and pulse broadening tomography may
urements, the downhole sonic probe is considered to give therefore correlate better with variable geology or variable
a smaller-scale, and usually higher velocity magnitudes structure when in situ stresses are very high, or where
than the averaged cross-hole result. However, the small- deep mining is involved.
scale excavation damage zone (EDZ) that may accom-
pany a borehole in incompetent rock, may be the reason
Chapter 5 Relationships between
for sometimes measuring lower local velocities at the
rock quality, depth and
small scale. Cross-hole measurement generally shows a
seismic velocity
smoothed, average behaviour. While general trends are
similar, the details between the cross-hole and sonic logs
5.1
differ, due to the change of scale and sampling location.
The simplicity of first arrival P-wave velocities from
4.5 shallow refraction seismic, and the easy access for core
drilling have stimulated various correlations between
When seismic velocity tomography is performed at dif- Vp, RQD, joint frequency and the rock mass quality
ferent scales around loaded rock samples, across loaded Q-value. For hard, low porosity rock a useful rule-of-thumb
564 Conclusions

is that Vp  3.5  log Q km/s, where Q is the rock both RQD and Qrock typically increase rapidly in the first
mass quality (range of Q  0.001 to 1000). tens of meters, making a reliable depth correction prob-
lematic, since the three variables quality, depth, and veloc-
ity are often all changing at once. Sensitivity of velocity
5.2 to stress is greatest in these first few MPa of stress
increase. The velocity-depth relation is non-linear, and
When rock has a uniaxial strength ( c) significantly has a steep gradient. Velocity increase with depth can
lower (or higher) than the nominal 100 MPa for hard occur in harder rock masses, even when rock mass qual-
rocks, the Q-value needs to be modified to Qc  Q  ity (Q-value, RQD, or joint spacing) remains constant.
c/100 in the above Vp–Q relationship. When rock
matrix porosity (n) is significantly more than the nominal
1% for hard rocks, Vp must be corrected by a porosity- 5.6
related reduction in velocity. When depth is significantly
greater than the nominal 25 m for shallow refraction seis- A measured 20 metres deep profile of weathered granites,
mic, a depth or stress related increase in velocity occurs, showing improvement with depth of all the indices of
which can also be estimated in the case of rock masses quality (hardness, RQD, density, etc.), demonstrated a
dominated by ‘soft’ porosity, or jointing. A non-linear very large increase in Vp from 1.0 to 4.5 km/s. The asso-
correlation of Vp with static deformation modulus has ciated velocity-depth gradient of 175 s1 emphasises the
also been developed, which can be utilised both with, or potential value of rock quality description. Velocity-depth
without, knowledge of the rock mass quality Q. gradients as high as 80 s1 in the upper 20 m, and as high
as 40 s1 over the first 50 m were recorded at a cavern site
in jointed gneiss with only slight weathering. Unusually,
5.3 the rock quality parameters: RQD, F m1 and Q, did
not improve at this site between 5 m and 60 m depth. A
Data from sites in chalk, chalk marl, sandstones, mud- maximum velocity rise of some 2 km/s (3.5 to 5.5 km/s)
stones, shales and tuff have been used to develop empiri- occurred in the depth range 10 to 60 metres, in which
cal Qc – velocity – depth corrections, which are non-linear horizontal stresses were interpreted from stress measure-
in the case of depth or stress level, but nearly linear in the ments to have increased by 2 MPa to 4 MPa, depending
case of porosity. on direction relative to h min and H max.

5.4
5.7
For near-surface measurements in harder rocks, the ratio
Fracture zone widths in the upper tens of meters tend
of P-wave velocities Vfield/Vlab., when squared, has often
to be larger if the velocity outside the fracture zone is
proved to be numerically close to the value of RQD (with
also low. Narrowest zones tend to have lowest internal
RQD expressed as a ratio rather than a percentage). RQD
velocities, and highest external velocities. There is com-
is defined as the % of core that has core sticks 10 cm
monly a reduction of the widths of low velocity zones
long. It applies to a given core run, or to selected struc-
with increased depth.
tural domains, or to specific rock types, or to cycles of
rock types, where there is some continuity. Since stress
can ‘acoustically close’ joints in weaker rock, joint fre-
5.8
quency as reflected in RQD may then prove to have little
correlation with the velocity. This is where other seismic
When apertures are less than approx. 0.04 mm (or
attributes than P-wave velocity become important.
40
m), the frequency of such fractures appears to have
little influence on the P-wave velocity. This is implied
5.5 by experiments with ‘line-samples’ in the form of multi-
ple-jointed columns of carefully machined rock cylinders.
Increasing depth will usually cause both the vertical Seismic surveys underground also suggest that apertures
and horizontal stress to increase. Besides stress increase, need to be considerably wider than 40
m, for fracture
Conclusions 565

or joint frequency (F m1) to influence the velocity to about 1 to 10 m, typical of the weathered zone. At
ratio Vjointed/Vintact. 0.25 MPa normal stress, equivalent to about 10 m depth,
conditions varying from no gouge to 2 mm of gouge,
may give S-wave velocity reductions of as much as 50%.
5.9

No reduction in P-wave velocity is predicted with as lit- 5.13


tle as 1 joint per meter. However, with 5, 10 or 20 joints
per meter, a hard crystalline rock with Vintact of 5.5 km/s, Careful experiments have demonstrated that shallow seis-
is predicted to show reductions to 4,100, 3,300 and mic refraction measurements that operate at low stress lev-
2,500 km/s respectively. Strong sensitivity of Vp to els are likely to be successful in distinguishing joint
stress level in the range 0.3 to 3 MPa is implied for the frequency and aperture. Amplitude measurements appear
case of filled discontinuities, including sensitivity to the to be much more desirable than velocity measurements at
total cumulative joint aperture and moisture condition. high stress levels, if joint frequencies and character are to
Vp is found to be proportional to n in the stress range be distinguished. This emphasises the value of attenuation
3–20 MPa, but drops rapidly when n  3 MPa. These measurement, or of seismic Q. Reduced amplitude and
experimental findings are similar to the empirical, field- increased attenuation occurs as joint roughness increases,
data based Qrock – VP – depth model. which fits with the picture of joint closure difficulties
when joints are rough. Shearing and dilation similarly
reduces amplitude and increases attenuation.
5.10

Compressive wave amplitude gives a more sensitive meas-


urement of the density of jointing than P-wave velocity. Chapter 6 Deformation moduli and
The amplitude ratio A/A0 (A0 for intact rock) shows seismic velocities
excellent sensitivity to the density of jointing. Apertures
less than 0.01 mm (10
m) apparently do not have influ- 6.1
ence on the wave propagation, even when the normal
stress is as low as 1–2 MPa. Physical apertures (E) of this It is well known from dam site investigations that the
size, in contrast to hydraulic apertures (e) which are dynamic modulus of elasticity (Edyn) that is calculated
smaller due to roughness effects, are probably rare in the from Vp, Vs and density exceeds the pseudo-static mod-
upper 20–30 metres of rock masses where refraction seis- ulus of elasticity (Ee) obtained from plate-load tests. The
mic is carried out, so this result is probably consistent latter is itself usually several times larger than the static
with experience in the field. modulus of deformation Ed, i.e. Edyn  Ee  Ed. In
poor quality rock masses, ratios of Edyn/Ed may be as
high as 10 to 20, due to the fundamentally different lev-
5.11 els of strain involved, but in massive hard rocks under
high stress the ratio will be quite close to unity. Ed is var-
Hydraulic apertures of about 10, 1.8 and 0.3
m are iously termed Estatic, D and M in the diverse literature.
implied with E 10
m, if joint roughnesses are respec-
tively 2.5 (quite smooth and nearly planar), 5 (near-
planar but some small undulations) and 10 (non-planar 6.2
with marked inclined asperities. The smallest of these
apertures would hardly be considered as ‘open’ joints. The differences between static and dynamic moduli are
attributed to the different strain amplitudes involved
(perhaps 103 and 106 respectively). The different
5.12 magnitudes of these moduli are also caused by the pres-
ence of pores, cracks and joints, which are measurably
Shear-wave velocity gives a very sensitive indication of the deformed in pseudo-static tests, but only very slightly
effect of gouge thickness at low stress levels, equivalent deformed by dynamic waves. When stresses are very
566 Conclusions

high and the pores, cracks and joints are almost closed, 6.6
the static and dynamic moduli are likely to be of almost
equal magnitude in the common direction of loading. Since seismic velocity also senses the effects of higher
joint stiffnesses, modulus can apparently be estimated
from the empirical relation M 10(Vp0.5)/3 GPa, even
without reference to the rock mass quality Q c which is
6.3 common to both. Where seismic ‘closure’ in relation to
Vp measurement occurs at shallower depth in weak
The modulus of elasticity (Ee) is traditionally obtained rock, the above relation may cease to track the assumed,
from the gradient of the unloading curves, which tend to continued increase in moduli. One may then revert to
have elastic character due to the frequently nearly closed the Qc value and apply the appropriate porosity ()
state of the stress-deformation loops. When unloading and depth () adjustments, in order to predict the con-
from higher stress levels, as when back-analysing shaft or tinued, assumed modulus increase.
cavern deformation, higher values of both Ed and Ee are
indicated. The rock mass becomes stiffer both in load-
ing and unloading, due to the increased degree of joint 6.7
closure as depth or stress increases.
Deformability data have been measured at dam sites in
weathered or soft rocks with uniaxial strengths ranging
6.4 as low as 2 MPa, and at sites with extremely hard rocks
with uniaxial strengths as high as 300 MPa. These rela-
The total deformation measured at the highest load level, tively shallow civil engineering sites may have in situ
after several load-unload cycles, is the usual basis for the P-wave velocities ranging from about 0.5 up to 5.5 km/s.
calculation of Ed. The applied stress level is usually based Corresponding ranges of Ee may be from about 0.5 to
on the size and type of dam, or other structure to be 80 GPa, and ranges of Ed may be from about 0.1 to
located on the particular rock foundation. Clearly, this 50 GPa. At intermediate rock strengths and intermedi-
limits the range of stress over which these parameters have ate velocities of 3 to 4 km/s, Ed and Ee values tend to be
typically been investigated. However, back-analysis of from about 5 to 10 GPa and 10 to 20 GPa respectively.
tunnels and caverns and deeper shaft deformations, and Porous rock, such as weaker limestones, results both in
higher pressure testing for large bridge foundations, has lower velocities and in lower moduli.
extended the near-surface range of stress.

6.8

6.5 The inequality Ee  Ed applies to the two rock engineer-


ing parameters that are actually in the ‘static’ loading,
Static deformation moduli (also termed Estatic, and M), ultra-low-frequency sphere of geophysics. The inequality
can be estimated from the non-linear relationship with Vp, of the moduli is due to mechanical hysteresis, and is not
or from the rock quality Q-value, c, n% and depth cor- related to another key inequality in geophysics, namely
relations developed in this book. In the undisturbed state, that the (static) joint stiffnesses become less than the
the consolidated state of the rock mass ensures a depth- inverse of the (dynamic) joint compliances, as rock mass
dependent deformation modulus, due to the tighter inter- quality deteriorates.
lock and higher stiffnesses of the rock joints. One should
be aware that most rock masses observed or tested, are
actually on part of a major unloading curve due to erosion, 6.9
or due to the rock excavation involved in preparing for the
test. In the case of rock masses with rougher, interlocked Plate loading tests taken to such high stress that rock mass
joints, lower joint stiffnesses may be registered, than if the failure occurs are unfortunately extremely rare. The meas-
jointed rock mass was loaded up without this prior urement of P-wave velocity at such sites may allow tenta-
unloading. tive extrapolation to other sites through a common rock
Conclusions 567

mass quality estimate. Such data are then the source of moduli Ed values as low as 5 to 10 GPa. The most likely
tentative rock mass strength ( c mass) estimation. The fol- explanation is that the large scale deformation modulus
lowing selected sets of (smoothed) data reflect the poten- test, as practised at dam sites (plate load, flatjack or occa-
tial linkages: Vp  2.3, 3.7, 4.0 km/s, Ed (or D)  1, 3, sionally pressure chamber) is nearly always registering an
15 GPa, c mass (uniaxial-loading)  4, 20, 50 MPa. excavation disturbed zone (EDZ) in a loading direction
parallel to the (unloaded) radial stress ( r) direction.
In contrast, the velocity measurement may be averaging
6.10 velocities over a larger volume, and may be recording
velocities parallel to the (tangential stress) direction,
It has been recognised in soil engineering that strain lev- or perhaps axially along the test adit wall. The tangential
els associated with normal foundation designs are rather stress is a much higher, maximum local principal stress,
small, for example, 0.01 to 0.1%, and therefore stiff- compared to the minimum radial stress, which approaches
nesses may be successfully described by the correlations zero, due to the effect of excavating the test adit. The
obtained from in situ seismic measurements. Such meas- axial direction would have intermediate stress levels, and
urements have the advantage of registering the stiffness presumably intermediate velocities as a result.
of the ground at in situ stress levels and in the undis-
turbed condition. When a tunnel or test adit is con-
structed at considerable depth in rock, the excavation Chapter 7 Excavation disturbed
disturbed zone (EDZ) effect will alter the above condi- zones and their seismic
tions in a complex way, to a degree that depends on rock properties
quality and the care with which the excavation of the
access adit and block test site has been performed. 7.1

Excavation disturbed zones (EDZ) caused by tunnel, slope


or open excavation may show up to several km/s reduc-
6.11 tion in P-wave velocity due to the combined effects of
radial stress relief, blasting or excavation damage causing
The three dynamic moduli (Young’s, shear, bulk), can the- fracturing. Improved drainage and drying-out of the
oretically be estimated from Vp measurement alone, if the near-field rock mass will also contribute to the reduction
dynamic Poisson’s ratio is estimated, rather than derived in velocity. Time effects, reducing Vp even more, may
from Vp and Vs. This however can cause significant inac- result from deformation or deterioration caused by inad-
curacies, and such values given in the literature should be equate rock mass reinforcement or surface protection.
treated with caution. It is normal to register lower
dynamic Poisson’s ratios in the case of higher velocity lab-
oratory data, and the opposite trend for the lower velocity 7.2
field data. High dynamic Poisson’s ratios are a sign of the
influence of jointing. In shear zones and faulted rock, very Cross-hole seismic monitoring of a ship lock excavation,
high values (0.4) of dynamic Poisson’s ratio are com- which reached a depth of more than 20 metres, caused a
mon. Static values of ‘Poisson’s ratio’ (mass-expansion- 200–300% reduction in velocity, a 75 to 85% reduction
coefficient) 0.5 may be measured when loading jointed in deformation modulus, and a 1 to 20 times increase in
rock towards shear failure. At hard rock sites, typical in (back-calculated) ‘joint voids’. The combined effect of
situ values of the three dynamic moduli Edyn, k and
loosening caused by blasting, stress relief, presumed inad-
might be 65, 50 and 25 GPa when the average P-wave equate slope reinforcement, and a one-year delay between
velocity is as high as 5.5 km/s. two of the monitoring stages, were responsible for the eas-
ily monitored degradation in properties.

6.12 7.3

Quite high P-wave velocities, such as 4.5 to 5.5 km/s are A larger scale version of stress relief at a gorge, which
sometimes reported together with static deformation cuts through massive, bedded limestones, had an outer
568 Conclusions

layer of 5 m thickness with a velocity of 2.9 km/s, background stresses are more isotropic. Illustrative mean
while deeper into the walls of the gorge, the velocity was results for the three zones may be Vp  3.5, 5.5 and
5.5 km/s. The potentially greater effect of weathering in 4.5 km/s respectively. Least effects on Vp tend to be seen
nature may be thwarted by lack of multiple joint direc- when tunnelling in saturated massive rock at greater (but
tions and permeability enhancement. Slope ‘EDZ’ may pre stress-fracturing) depths, and greatest effects on Vp are
exceed the EDZ associated with tunnelling, due to the seen in drained or dry jointed rock at shallower depth.
longer and more effective influence of surface weather- The hazardous stress fractured zones of tangentially-
ing, including frost damage. strained and radially-loosened or dynamically loosening
rock, would show lowest velocities if measurements could
be achieved.
7.4

Seismic attenuation can be used to evaluate the efficiency 7.7


of rock excavation by blasting, by quantifying the degree
of brokenness of the rock before and after blasting. Mining of excavations in bedded salt has shown greatest
Certain frequencies may be attenuated more than others effect on the attenuation of P-waves. Mining induced
due to the effect of fracture size and fracture-induced radial stress relief and increased tangential stress tend to
voids. Attenuation will tend to be larger and more irregu- cause under-saturation due to dilation of the salt. A reduc-
lar at shallow depths, where existing joints and new frac- tion in P-wave amplitude, which may increase with time
tures open more easily. Attenuation is more regular and after excavation, may give a stronger indicator of the EDZ
more limited at greater depths, where confinement lim- than a minor reduction in P-wave velocity.
its fracture and joint opening. Such EDZ effects may be
accompanied by rotations of both the attenuation
anisotropy axes and the velocity anisotropy axes, as a
result both of disturbance to pre-existing joint patterns 7.8
and dissipation of prior stress anisotropy.
The effect of radial stress release close to a tunnel excava-
tion in jointed rock, generally increases the joint aper-
7.5 tures, which can actually be expressed as a void ratio.
A time-average equation utilising the P-wave velocity
The removal of stressed rock and its usual replacement by through air or water-filled joint voids, together with the
air at atmospheric pressure when tunnelling, results in a P-wave velocity of intact rock, can be utilised to interpret
radial stress ( r) that approaches zero at the excavation depth-dependent downhole or up-hole P-wave velocities.
walls. The tangential stress ( ) usually increases in dia- The result is explained in terms of an increased joint-
metrically opposite sectors, but may also incorporate dia- aperture related void-ratio, with reduced depth of meas-
metrically opposite sectors of negative min if the urement from the excavation walls.
far-field stress anisotropy is 1 : 3. Joint set orientations
relative to an anisotropic stress, adverse rock strength/
stress ratios, and the disturbance caused by the excavation 7.9
method (blasting, boring, or line-drilling), may each have
strong influence on the physical EDZ, and on the result- Cross-hole seismic measurements performed in a colum-
ing seismic EDZ. nar-jointed basaltic rock mass at BWIP, Washington, USA
were made between four horizontal boreholes drilled 12
metres into the wall of a drill-and-blasted underground
7.6 opening at 46 m depth. The basalt columns were regular
but sinuous, 0.15 to 0.36 m in thickness, dipping 70 to
Three velocity zones may be visualized around a tunnel: a 90°, with frequent low angle, discontinuous cross-joint-
1 m or less loosened zone with lowest velocities, a stress ing. The horizontal and (to a lesser extent) the diagonal
bearing ring with highest tangential stresses and velocities, seismic measurement paths crossed the more open colum-
and an uninfluenced zone with reduced velocity where the nar joints, and these features clearly opened most as a
Conclusions 569

result of excavation, giving the strongest reductions in AE events clustered both where tangential stresses were
velocity of 55% to 65%. The seismic quality factor Qseis highest and where seismic velocity (Vp) gradients were
increased in the vertical direction, in the same direction as steepest. The acoustic emission results confirmed that
the highest velocities recorded where there was least EDZ rock failure was initiating just inside the tunnel wall,
effect. Qseis in the horizontal direction remained low (5 to orthogonal to the 1 direction. Relatively decreased veloc-
8), in the same direction that showed the maximum ities were seen in the two regions that were under tensile
EDZ-reduced velocity, with VP declining from 5.5 or tangential stress. Calculated P-wave velocities showed
6.0 km/s to 3.5 km/s at the tunnel wall. quite strong anisotropy in the massive granite, caused
by the principal stress anisotropy given in the previous
paragraph.
7.10

The underground research laboratory (URL) in 7.13


Manitoba, Canada was the site of numerous geophysical
studies. The dominance of massive, unjointed, highly An experimental tunnel sealing experiment at the URL
stressed granite resulted in particular focus on stress- utilised AE for interpreting reductions in average P-
related EDZ, with down-hole sonic logging and acoustic wave and S-wave velocities, in the highly stressed zones
emission monitoring, together with parallel laboratory that were caused by post-excavation of larger diameter
tests and numerical modelling studies. At the smallest ‘dog-collars’ or bulkheads for concrete and bentonite
scale, the state of micro-cracking in selected core samples sealing. 3D coverage of an outer volume surrounding
from radial boreholes drilled in the walls of a drift where this tunnel sealing experiment, and two higher resolu-
excavation was either by normal blasting or by smooth tion AE arrays for recording in a 10 m  10 m  10 m
blasting, indicated Vp reductions of about 1 to 1.5 km/s volume around the collars, allowed AE-based monitor-
in the outer 0.8 m of the normally blasted tunnel, and ing of temporal changes in Vp and Vs. These were used
reductions of about 0.5 km/s in the outer 0.5 m of the to estimate the theoretical change in crack density (c)
smooth blasted excavation. and saturation (s) along any particular ray path.

7.11 7.14

The effects of highly anisotropic, sub-horizontal stresses At the Äspö hard rock laboratory (HRL) in Sweden, seis-
at URL were studied in a unique test tunnel excavated by mic tomography investigations were performed to com-
line drilling and reaming, followed by mechanical break- pare the depth of excavation damage zones in immediately
out to avoid blast damage. Principal stresses of approxi- adjacent drill-and-blast and TBM sections of tunnel.
mately 60, 45 and 15 MPa, caused classic ‘break-out’ Principal stresses at the ZEDEX (zone of excavation dis-
resembling that in a borehole. The isotropic-elastic theo- turbance experiment) were approximately 32, 17 and
retical tangential stresses of 165 MPa (3 1  3) at 11 10 MPa. There was only a small EDZ effect on Vp and Vs,
o’clock in the roof and 5 o’clock in the floor (with due to the high stresses and the partly discontinuous joint-
3 3 – 1  15 MPa in the side walls), caused promi- ing. P-wave velocities were mostly in the region of 6.0 to
nent V-shaped notches of rock failure. The stress-related 6.3 km/s, with small reductions of velocity in a thin ‘skin’
disturbance was measured directly in 1 m deep boreholes next to the walls. The reductions of velocity were recorded
using a micro-velocity probe with 10 cm separation of in the first 0.25 m into the TBM tunnel walls, and up to
the transducers. 1 metre into the walls of the drill-and-blasted tunnels.

7.12 7.15

Acoustic emission (AE) was recorded during careful EDZ effects around the tunnel were only just detectable
(mine-by) extension of the 420 m deep test tunnel. The by seismic velocity, due to the good quality of the rock
570 Conclusions

(Q-value  22–24), and due to stresses that were high 7.18


enough to have acoustically closed the joints, but not
high enough to have caused excavation induced micro- The average joint frequency in the test area, analysed
cracking, as at the URL in Canada. The rock quality Qc- from 224 m of core, was 8.3/m. An elastic continuum
Vp-M model predicted a P-wave velocity range from 5.8 analysis conducted prior to the test had suggested larger
to 6.2 km/s, using an equivalent depth range of 400 to stresses and local displacements than were actually meas-
1200 metres, relevant to the principal stress range. The ured, presumably due to the thermal compliance of the
intact rock laboratory E-modulus was 69 GPa. The equa- joints, as noted in Part II concerning a fully coupled
tion relating mean deformation modulus M with Vp, HTM heated block test. This thermal compliance effect
suggested deformation moduli ranging from 58 to was also experienced in a heated-mine-by experiment
79 GPa, for the predicted velocity range of 5.8 to in the Climax Mine, in the USA, where deformation
6.2 km/s. Calculated dynamic moduli around the drill- monitoring also failed to match predictions, since no
and-blast drift ranged from 76 to 79 GPa. The particular thermal joint compliance was modelled.
M  Edyn situation at this site implies ‘acoustically closed’
jointing. Chapter 8 Seismic measurements
for tunnelling

7.16 8.1

A drift in the Stripa mine in Sweden, used for borehole Seismic velocity measurements performed from the sur-
heater tests, showed increased seismic velocities between face prior to tunnelling need careful interpretation due
drained monitoring holes in the jointed quartz mon- to the common experiences of enhanced weathering,
zonite, during heating of the rock mass. The initial enhanced fault width and more open jointing experienced
increase in velocity with temperature was linear and var- near the surface. Anisotropic stresses, and anisotropic fab-
ied from 2 to 4 m/s/°C. This was the presumed result ric or structure or bedding, plus hidden lower velocity
of thermally induced joint and micro-crack closure. layers, add to the potential pitfalls when attempting to
Increased P-wave velocities were recorded in directions anticipate rock mass conditions at tunnel depth. The
consistent with a thermally-induced stress increase. An degree of saturation where measurements were made,
initial stress anisotropy giving low stress in the same direc- compared to tunnel depth conditions is also important.
tion as the subsequent thermal stress increase tended to
enhance the Vp response. Greatest response was seen when
8.2
degrees of saturation were also low.
Seismic refraction profiles with appropriate azimuth,
and borehole dip and dip-directions that also take due
7.17 regard of structure and stress anisotropy can improve
the quality of the sub-surface investigation. The use of
A long period of cooling (350 days) generally returned only vertical boreholes drilled where there is predomi-
seismic velocities to values lower than before the heating, nantly vertical structure is a guaranteed way to obtain
suggesting permanent changes, probably connected with poor sub-surface information, if this is only to be based
the hysteresis effect of joint closure and opening. Upon on core and borehole wall inspection. The application
cooling, the less rough, interlocked joints may have of geophysics helps to recover some of the lost informa-
‘sprung-open’ more than their closed neighbours, to avoid tion from representative core samples, especially if high-
tensile stress development. This would cause a reduction frequency cross-hole surveys are performed.
in seismic velocity if the open joint or joints crossed the
path of the seismic array. A velocity anomaly at about
3 m depth was smoothed-out by the heating but returned 8.3
when the rock was cooled. A significant quantity of water
expelled during the heating signified a general closing The common experience of improved rock quality and
of the joints. increased Vp at depth may be checked at intervals by
Conclusions 571

appropriate core-drilling. In the local absence of core, a for. This may mask the actual presence of a serious fault
sensible interpretation of the usual effect of depth on Vp that could better be identified by shear waves or atten-
values can be made using QVp correlation, where Q is uation measurement.
the rock mass quality. In very general terms at hard rock
sites, a ten-fold improvement of Q-value, combined with
8.7
a 1 km/s additional depth effect on VP might be expected
at a 50 m deep tunnel. So a near-surface Q-value of 1.0
An early example of the use of geophysical surveys in tun-
interpreted from shallow seismic refraction, where VP was
nels was the Straight Creek pilot bore of 4.0 m diameter,
3.5 km/s, might see VP increased to 5.5 km/s at 50 m
driven in the 1960’s at 200–500 m depth through granite,
depth, but with the Q-value equal to 10 at tunnel depth,
diorite, gneiss, migmatite and schist, under the continen-
rather than Q 100 as implied by a nearer-the-surface
tal divide in Colorado, USA. Deep layer velocities were
VP of 5.5 km/s.
measured at five seismic spreads: 5.18 km/s, 5.1 km/s,
4.8–6.1 km/s, 4.2 km/s and 6.0 km/s. Much lower shallow
8.4 layer velocities, representing loosening effects were respec-
tively 3.0 km/s, 2.3–2.7 km/s, 2.3–3.1 km/s, 1.3–1.6 km/s
There is evidence to suggest that not only rock over- (worst case, class 5) and 2.3 km/s. These extremely low
burden depth, but also water depths are important, in EDZ velocities proved, in retrospect, to be more related to
the case of undrained sub-aqueous tunnels that prove the insurmountable problems in the 12 m main bore,
to be completely dry, due to low permeability over-lying which took several year to complete. There was possibly
rock masses such as phyllites. In such cases the total insufficient appreciation of the effect of stress on the seis-
stress caused by rock load and water load may give a mic velocities at that time. There is also the possibility of
false impression of rock quality at tunnel depth, to the adverse interaction when there are twin tunnel tubes, with
extent of 1 to 2 km/s. ‘plastic zone’ overlap (or log-spiral shear-zone overlap) in
faulted or weak rock zones, a problem of relevance when
assessing risk in too-close twin-bore TBM tunnelling,
8.5 where conditions are unfavourable.
Seismic velocity measurements performed while tun-
nelling have been shown to correlate in approximate 8.8
terms with speed of tunnelling, support needs, and tun-
nel cost. Probe drilling and sonic logging of the holes Comprehensive geological and velocity classification of
can be utilised for ahead-of-the-event information on rock conditions at numerous rail tunnels in Japan from
rock and hydrogeological conditions. If multiple probe the 1970’s demonstrate the value of velocity-tunnel-
holes are drilled, then cross-hole tomography can be support correlation. Recommendations were based on
performed to add to the tunnel contractor’s prepared- 70 case records from 30 m2 and 60 m2 tunnels. For rock-
ness for difficult zones. This has resulted in the choice type and geological Classes 1 to 6, with their associated
of ground-freezing for safer penetration of dangerous high to low P-wave velocities, tunnel support loads ranged
and/or environmentally sensitive sub-urban fault zones. from 1 to 30 tons/m2, the spacing of the steel arch sup-
port ranged from 1.5 to 0.75 m, and the final concrete
thickness ranged from 0.3 to 0.9 m. Today, we could
8.6 place an equivalent velocity scale below the rock mass Q-
value, which is used for recommending shotcrete thick-
Velocity measurements should be made sufficiently far ness and bolt spacing.
ahead of the tunnel face, at least two to three diameters,
such that additional stress concentration effects are
avoided. Interpretation of local effective stress condi- 8.9
tions will be needed when evaluating results of the
velocity measurements, using VpQ correlation. A When attempting to add a rock mass Q-value scale
‘false’ rock quality, due to a higher P-wave velocity may to such velocity-rock-support recommendations, the
be interpreted, if depth of measurement is not accounted geological part of the classification is particularly
572 Conclusions

important, in view of the porosity, strength and depth empirical model can be applied to make these assess-
corrections to the VP  Qc empirical model. Young, ments, where Qc  Q  c/100.
weak, porous rocks may have RQD  0, and incur
unfavourable SRF values due to adverse strength to stress
ratios in the tunnelling situation. These two factors may 8.12
alter the correlation of support-type-and-degree with the
velocity, unless the velocity is EDZ-based, rather than that Despite the above pitfalls concerning the meaning of
obtained from the fully confined (pre-tunnel) situation. velocity when different porosities and rock strengths are
involved, tunnel refraction seismic and tunnel boring
machine (TBM) progress indicate a quite linear inverse
relationship between the penetration rate (PR in m/hr,
8.10
with uninterrupted boring), and P-wave velocity. This
is also partly matched by an inverse PR and Schmidt-
Sub-sea tunnelling experiences, with velocity-rock sup-
hammer rebound trend. Respective values might be as fol-
port correlation, have revealed several cases of medium
lows: VP  4 km/s, 2 km/s: PR  2 m/hr, 4 m/hr:
and low-velocity zones actually creating greater tunnelling
Schmidt-hammer rebound  40, 20. (Inverse correlation
difficulties than the velocities would suggest, implying an
to possible Q-values of 30 and 0.3 are also implied, as in
artificially elevated velocity in relation to rock quality. The
the QTBM model). Harder, higher velocity rock signifies
tunnelling situation changes the stress level in relation to
tougher boring (i.e. lower PR), but the weekly mean
the pre-tunnelling confining stress. Based on sub-sea tun-
advance rates (AR) might be 1 m/hr and 0.5 m/hr respec-
nelling experiences, Norwegian studies have shown sup-
tively, due to the reduced tunnel support needs. While
port costs rising from 50% to at least 75% of total costs,
tunnel support quantities are roughly proportional to
when the P-wave velocity reduces from 5.5 to 4.5 km/s.
log10 Q, PR may be inversely proportional to log10 Q.
It is likely that these are ‘depth-enhanced’ velocities, since
uncritically applied Q-values would be 100 (no support)
and 10 (light support) respectively, which are unrealistic
8.13
when referring to significant support costs.
Sonic logging of probe holes, made by fast percussion
drilling at from 2 to even 5 m per minute, can give
8.11 advance information on seismic velocity if the holes are
sonically logged. This can be used for subsequent rock
High velocities of typically 5 to 6 km/s may dominate quality class and tunnel support class estimation. The cho-
in refraction seismic studies at relatively unweathered sen excavation mode, and support components such as
hard rock sites. The much smaller number of tectonic bolted steel arches, rock bolts and mesh, or shotcrete, can
zones (shear zones, faults), dykes and joint swarms with then be immediately available, and applied with appropri-
velocities from about 2.5 to 3.5 km/s are what cause the ate timing, behind the advancing TBM tunnel face.
construction problems, especially when high inflows of
water occur. In weaker, porous rock, velocities in the
range 2.5 to 3.5 km/s may signify excellent stability. 8.14
The contrast from back-ground velocities is clearly the
measure of the stability problem, not the velocity per A seismic velocity probe ahead of a tunnel will not see
se. The hard rock velocities may signify rock quality the difference between a TBM tunnel and the drill-
Q-values of 30–300 (very/extremely good), and and-blast tunnel. However, if refraction seismic mea-
0.1–1.0 (very poor) respectively, clearly reflecting the surements are performed along the wall of a TBM
reduction of quality, and need for local, heavy tunnel tunnel, the values of Vp obtained may tend to be higher
support. The weak, porous rock velocities could signify than in the equivalent drill-and-blasted tunnel for at
Q-values as high as 10–100 at nominal 25 m depth, if least three reasons. There is a reduced level and depth of
corrections for 10% matrix porosity and UCS of damage in the wall of a TBM tunnel. Higher tangential
10 MPa were included in the assessment. The VP  Qc stresses are acting closer to the TBM tunnel wall. There
Conclusions 573

will be a tendency for lower permeability and less Chapter 9 Relationships between
drainage around the TBM tunnel, which, for reasons of Vp, Lugeon value,
more complete saturation might also increase the seis- permeability, and grouting
mic velocity. On the other hand, there may be a reduced in jointed rock
value of effective stress as a result of the same reduction
in permeability. 9.1

Rock masses containing voids in the form of porosity,


8.15 joints or damage zones will generally have enhanced per-
meability and reduced seismic velocity. Measurements at
The reflection method of HSP or TSP (horizontal or dam sites and at tunnels have indicated inverse correla-
tunnel seismic profiling) with both source and receiver in tion between P-wave velocity and permeability, in partic-
the tunnel, is capable of locating seismic reflectors ahead ular with the relatively high pressure Lugeon injection
of a tunnel face. However, reflector images are not related test, which may locally deform the void space created by
to rock quality directly, but to implied change of quality. joints or fractures. Because the rock mass quality or
It is difficult to determine if the rock quality will get bet- Q-value also correlates with the deformability, a very sim-
ter or worse at a given reflector, and there may be inac- ple lower bound inverse relation between the Lugeon-
curacies of reflector location due to the unknown actual value (L  107 m/s) and the Q-value is indicated. (Q 
velocity field. 1/L). This is not relevant however, when joint-sealing clay
is present, causing both Q and permeability to reduce
together, instead of inversely.
8.16

A logical extension of conventional high-resolution 9.2


surface refraction seismic, is the application of sources and
receivers at, and close behind the tunnel face, with Lower Q-values and higher Lugeon values will tend to be
receivers at the surface, if this is not too distant. GPS measured when holes are oriented to cross major struc-
clocks are needed to synchronise the sources within ture. The opposite occurs when paralleling major struc-
the tunnel and the receivers at the surface. The use of in- ture, and failing to cross the dominant jointing. The latter
tunnel refraction-based estimates of velocity distributions is typical for vertical wells prior to deviation through
ahead of the face, can improve the accuracy of reflector seismically identified oriented structure. Holes should be
positions. With the necessary velocity distribution ahead drilled in the ‘slow’ direction for registering highest per-
of the face, the rock mass quality can be characterized both meability. Clay sealing of joints may compromise these
up to and beyond the now better-located reflectors. simplified assumptions.

9.3
8.17
Since sub-vertical jointing may dominate in the same way
Velocities measured at depths of hundreds of meters or that horizontal stress anisotropy may dominate, the per-
more, using in-tunnel refraction or cross-hole tomogra- meability anisotropy will tend to be related to azimuth. At
phy, may bear little resemblance to the major tunnelling great depth, this model may be compromised by joint
difficulties sometimes experienced when tunnelling. Face closure, and a shear mechanism may be needed to explain
collapse in a Vp  4 km/s rock mass is ‘illogical’, without the maintenance of (anisotropic) permeability.
allowing for the depth or stress effect that may mask, in
velocity terms, the true low quality. A 300 m overburden
at such a collapse location, suggests a near-surface Vp of 9.4
about 2.5 km/s, using the Q-Vp-depth model. This would
be relevant to a serious fault zone, or extremely poor rock, The void space in a rock mass that is responsible for great-
and therefore more consistent with eventual collapse. est permeability, specifically the joints and any outwash
574 Conclusions

channels caused by weathering or soluble minerals, can showed a very broad transmissivity range of 102 to
usually be injected with cement, micro-cements, or ultra- 1014 m2/s. Corresponding P-wave velocities ranged
fine cements. Seismic measurements, principally in dam from 3.2 to 5.5 km/s, Vp/Vs from 2.45 to 1.72,
foundations, show remarkable increases in P-wave veloci- dynamic E-modulus from 0 to 60 GPa, and dynamic
ties as a result of successful grouting, when permeabilities Poisson’s ratio from 0.41 to 0.28.
are also favourably reduced. Part of the increase in veloc-
ity and resulting reduction in deformability may be due
to the post-stressing effect of high pressure grouting,
9.8
which also may set in a stressed state.
At the Chinnor Tunnel in chalk marl in southern
England, very low seismic velocities in the range 0.6 to
9.5 1.0 km/s were registered for badly fractured/jointed areas
of the chalk marl. Quoted permeabilities were 104 to
Most effective grouting can be expected in the joint set 106 m/s in these areas. Assuming that 1 Lugeon 
having greatest permeability and least magnitude of the 107 m/s then the very high Lugeon values of 1000 to 10
effective normal stress. This set will thereby show great- imply Qc values of 0.001 to 0.1. These low Qc values can
est increase in seismic velocity as a result of the grouting. be converted to ‘tunnel support’ Q values of 0.02 to 2
A rotation of the permeability tensor to a lesser magni- (‘extremely poor’ to ‘fair’), by assuming a mean c value of
tude can also be expected as a result of the grouting, and 5 MPa for the chalk marl. This range of Q-values is in line
velocity anisotropy is likely to be reduced. Several of with expectations for the heavily jointed rock mass at
these aspects have been demonstrated using multiple Chinnor.
borehole 3D permeability testing in Brazil.

9.9
9.6
It has been confirmed from comparison with data from
The inter-related physical nature of rock mass quality carefully documented block tests performed in the USA,
(Q), deformation modulus (M), seismic velocity (Vp) that the parameters Q, Vp, M and L are inter-related, and
and permeability (L) means that rock masses can be that the inclusion of the Lugeon value in this inter-rela-
represented by type curves in nomograms linking these tion is justified, if care is taken to eliminate irrelevant
parameters. Although only approximate, such type curves non-deforming, channel flow cases, and to eliminate cases
serve to distinguish typical properties of massive rock, where clay sealing of the joints is occurring. Depth or
jointed rock, porous jointed rock, and fault zones. The stress level, also plays an important role in these mutual
mass properties of each will also be affected by depth inter-relationships.
and by matrix properties such as uniaxial compressive
strength and porosity.
9.10

9.7 Using an analogue material for heavily jointed rock,


namely coal, one can see great sensitivity between veloc-
Extensive work in marl formations in Switzerland, indi- ity, stress level and permeability, which will also be present
cated quite strong relations between selected seismic in jointed rock masses at large scale, when in situ effective
parameters and transmissivity measurements in five stress states are altered by large scale pumping or injection
deep boreholes at Wellenberg, a potential nuclear waste experiments. Extensive data sets for numerous coal sam-
repository site. Good correlations with transmissivity ples show the three key behaviour modes: permeability
were obtained with Vp, Vs, Vp/Vs, dynamic shear (
) reduction with increased stress, permeability reduction
and E-moduli, and dynamic Poisson’s ratio. The bore- with increased velocity, and velocity increase with
hole depths ranged from about 400 m to 1800 m, and increased stress. Greatest changes in Vp and permeability
included faulted and brecciated rock. The measurements occur at the lowest stress levels and at the lowest velocities,
Conclusions 575

just as found in rock masses, due to the greater sensitivity 9.14


to acoustic coupling across the cleats in the coal.
Dam foundation damage in the form of shearing on folia-
tion planes, with resultant P-wave velocities as low as
9.11 0.5 km/s, was a starting point for systematic cement grout-
ing, that saw Vp locally increased to between 2 and
Researchers working at nuclear waste related rock labo- 6 km/s. The remaining low velocities of only 2 to 3 km/s
ratories such as Stripa, Äspö, and Grimsel have utilised correlated with residual permeability, which was elimi-
seismic and radar tomography to characterise fractured nated with further grouting, and consequently increased
zones and fault zones. Their studies have generally helped velocities. A strong velocity-depth effect was noted, related
to explain why these relatively small volumes of fractured both to rock quality improvement at depth, and to a
or heavily jointed rock are responsible for such large post-stressing effect from the increasingly confined grout
percentages of the total flow of water. Radar attenuation at greater depth. New rounds of grouting at old dams has
difference tomograms have been found more reliable in also demonstrated increased Vp and reduced permeabil-
locating brine in tracer tests, than slowness tomograms. ity. Largest grout take correspond to the locations where
the largest increases in Vp are registered, following the
grouting.
9.12

Radar and seismic signals are sensitive to different physi- 9.15


cal parameters (electro-magnetic wave conductivity, and
mechanical stiffness respectively). The respective tomo- It was noted in the 1970’s, in the 220 m high Mratinje
grams therefore highlight different features of the rock dam foundation, that effective consolidation grouting
mass. Radar may delineate permeable zones caused by could be performed when Vp was in the range 2.5 to
pore space or by joint apertures, in slightly different loca- 3.5 km/s. According to the Q-Vp-L correlations, this may
tions to the low seismic velocity zones associated with the correspond to Q  0.1 to 1.0 and 10 to 1.0 Lugeon.
clay-filled sections. The one will usually lie parallel to the Velocities above 4.0 km/s (implying Q  3 and K 
other, since higher permeability may be associated with 0.3 Lugeon) could not be improved upon by the grout-
the heavily jointed zones that are often found in the walls ing done at that time. Such results emphasise the rea-
of faults. sons for combined use of high injection pressures of 5
to 10 MPa, and use of micro or ultrafine cements and
microsilica, in today’s pre-grouting ahead of tunnels.
9.13 Systematic pre-injection where average Lugeon values
are as low as 0.1 (implying VP of 4.5 km/s), may not be
Low resistivity generally correlates with zones of increased successful without resort to high pressure and/or use of
water content and frequently with higher permeability. At the finest cement types.
a site in South Korea, a series of boreholes in weathered
granites were Q-logged and later compared with resistiv-
ity tomograms in a ‘blind’ test. Sections of the core with 9.16
increased joint frequency (low RQD, high Jn) did not
always correlate with low resistivity and vice versa. The A general, approximate relationship between Lugeon
two parameters that did show a consistent correlation value, Q-value, and measured velocity, has been com-
with low resistivity were low values of Jw estimated from pared with case records showing reduced Lugeon values
iron staining or apparent aperture, and the high values of and monitored velocity increases, as a result of grouting.
Ja due to sand or silt fillings, and due to clay-fillings. The These changes also represent potential physical effects of
latter gives ‘anomalously’ low resistivity due to the ionic grouting on individual Q-parameters. Conservative
effects of the clay. Water content and permeability are assumptions of individually improved Q-parameters, like
clearly lower in such discontinuities than in those that are increased effective RQD, reduced number of effective
sand or silt filled. This represents a potential source of joint sets, and successful grouting of the least favourable
error in judging the meaning of low resistivity zones. joint set, combine to suggest significantly improved
576 Conclusions

effective Q-values. Reduced tunnel deformation, reduced to the relaxed or unrelaxed state caused by low or high
support needs, increased modulus of deformation, and frequency.
increased seismic velocity are each implied, and seem to
be supported by the experience of trouble-free tunnel
advance following effective, high pressure pre-grouting. 10.3

The short period of each wave cycle, and the relatively low
PART II levels of dynamic stress, are generally assumed to cause
sub-micron size reversible micro deformations and flows.
Chapter 10 Seismic quality Q and There are those who have denied the early assumption of
attenuation at many friction as an intrinsic attenuation mechanism in the
scales earth, but this supposition was partly based on dynamic
sub micro-strain work with intact bars of dynamically
10.1 excited rock. Frictional losses in situ, due to the presence
of joints and fractures, plus their contribution to losses by
Attenuation can be simply defined as the loss of energy scattering, are now widely cited, and by various authors,
per cycle divided by the maximum energy per cycle in the when working outside the common intact-medium labo-
same rock volume. The wave amplitude versus frequency ratory limits. Important here is the fact that when the
diagram indicates that frequency determines the maxi- effect of single joints or aligned fractures are inverted
mum wave amplitude. The inverse of attenuation, the from seismic data, the respective contributions of the
seismic quality Q, shows high values when there is little dynamic stiffnesses of the joints to the wave amplitude
attenuation, and low values when rock conditions and and to shear-wave anisotropy, are found to have recognis-
near-surface location causes strong attenuation. The stan- able magnitudes and the same units (Pa/m) as when
dard definition: E/E 2/Q suggests a minimum the- working with pseudo-static loading tests. The latter obvi-
oretical value of seismic Q of about 6, but lower values are ously mobilize friction due to their contribution to fric-
reported, including possibly erroneous negative values. tional strength.
Several methods can be used to estimate seismic Q, which
can have the forms Q p and Q s when derived from P-wave
or S-wave spectral analysis, or Q c when derived from the 10.4
tail of an earthquake seismogram, termed the coda, to
sample the deep hypocentral region. Review of a large body of laboratory data for seismic
Q has indicated a close approximation to the pseudo-
static Young’s modulus (Estatic) and it’s increase with con-
10.2 fining stress and sample stiffness. This was noted when
substituting a non-linear seismic Q scale next to numer-
Knopoff’s paper ‘Q’ from 1964 with long lists of seismic ous sets of experimental data that showed only a linear
Q values for solids like steel, glass, lead and celluloid 1/Q attenuation scale. Laboratory studies frequently
(5000, 490, 36 and 7), suggest that relative stiffness may show seismic Q increasing from about 10–20 to about
be involved in some way. It was originally thought that 50–100, as confining pressure is increased from near zero
seismic Q was independent of frequency. However, rock to 10, 20, 40, or 70 MPa . Seismic Q magnitudes have
with microcracks, joints or fractures, and with different great similarity to the variation of Estatic with confining
fluid saturation levels, is now known to cause frequency- pressure, if seismic Q is expressed as if it were GPa.
dependent attenuation. Attenuation losses occur due to
wave scattering from heterogeneities like joints, faults and
rock boundaries, and due to intrinsic losses such as matrix 10.5
anelasticity, friction at grain boundaries, across microc-
racks and across joints and fractures, related heat loss, fluid Intact rock samples loaded to failure show consistently
‘squirt’ from fractures to microcracks to pores, and gas increasing seismic Q when measured in the loading direc-
pocket squeezing when partly saturated. The viscous fluid tion, as does Estatic, and the same samples show reducing
related losses are particularly dependent on frequency due values of seismic Q when measured in the perpendicular
Conclusions 577

(Poisson expansion) direction. Ultrasonic 0.1 to 1 MHz 10.8


seismic Q values estimated during normal-stress closure
tests on jointed samples, show Qp increasing from about With clay and weak rock at one extreme, the deformation
5 to 80, as normal stress is raised from about 3 to modulus range can extend down to 0.1 GPa, representing
70 MPa. At another extreme, cracking caused by a strongly absorbing inelastic matrix with UCS  1 MPa,
repeated freeze-thaw cycles in limestone, have demon- plus the near-surface attenuating effect of clay-filled dis-
strated Qseis steadily declining with increased cracking continuities. At the other extreme, with high matrix
from about 20–25 down to 10–12, with corresponding strength UCS  200 MPa, a depth up to 1 km and neg-
reductions in P-wave velocity. ligible jointing, Emass may reach 150 GPa, representing
very low attenuation.

10.6
10.9
In situ seismic interpretations of seismic Q have revealed
a certain tendency for higher seismic Q when the rock
In sedimentary, finely layered rock sequences, greater
mass quality Q-value was also expected to be higher.
sophistication in broad-band recording has demon-
These commonly used quality parameters are differenti-
strated a more complex Qseis-deformability response
ated as respectively Qseis and Qrock. The widely used Qrock
than the above, with seismic Qp indicating that different
value, varies from about 1 to 1000 for heavily jointed to
values are obtained from ultrasonic 500–900 kHz testing
massive almost joint free rock masses. It is composed of
of core (mean 27.0), sonic 8–24 kHz testing in boreholes
three ratios. The first ratio is RQD (% of core
(mean 10.4), cross-hole 200–2300 Hz testing (mean
pieces 100 mm length) divided by the number of joint
15.7) to VSP 30–280 Hz testing from surface to bore-
sets Jn. The second ratio Jr/Ja describes roughness and
hole (mean 31.3). The rock mass was a finely layered sat-
alteration, and gives the friction coefficient, which
urated sequence of limestones, sandstones, siltstones and
includes the effect of clay-filling. Finally the estimated
mudstones. The overall range of Qp from the different
water pressure (or tunnel inflow) and the stress/strength
frequencies and depths was approximately 5 to 40. This
ratio are evaluated. Qrock RQD/Jn  Jr/Ja  Jw/SRF.
remains in a potentially familiar range, if expressed in
The maximum range of Qrock is from 0.001 for fault
rock mechanics deformation modulus units of GPa.
zones to 1000 for massive joint-free rock masses. Qrock
provides quick estimates of seismic Vp, deformation
modulus and permeability, and also indicates tunnel sup-
port needs in rock engineering projects. 10.10

Coda waves from the tail of seismograms, typically in the


10.7 20 to 200 seconds time window, after the arrival of
P, S and surface waves, have traditionally been used to
The first pair of Qrock parameters RQD/Jn, representing analyse the seismic Q signatures of both local and dis-
‘relative block size’, have a numerical range of 200 to 0.5. tant earthquakes. A large volume of rock surrounding the
Even this has a certain similarity to the range of Qseis for earthquake source is believed to be sampled by the coda,
respectively joint-free to heavily-jointed rock masses. with the possibility to register changing rock properties
More important is the likeness of the rock mass (pseudo- or migrating source zones, due to the accumulation or
static) deformation modulus to Qseis, because Emass is read- dissipation of stress and resulting strains and possible
ily estimated from Qrock. The great majority of rock fluid movements. Coda Qc is frequency dependant, with
masses, from the surface to 1 km depth, from heavily low frequency 1 Hz content assumed primarily composed
jointed to virtually joint-free rock, have deformation mod- of surface waves scattered from shallow heterogeneities,
uli in the approximate range 1 to 150 GPa, as derived from while coda waves at 20 Hz are believed to be from back-
in situ testing and from back-analysis of deep tunnels and scattered body waves from deeper heterogeneities in the
shafts. This range closely resembles the typical ranges of high Qseis lithosphere. Thus many surface recorded data
in situ Qseis seen in numerous shallow 0 to 2 km deep seis- sets show Qseis lower than 100 at frequencies 2 Hz,
mic data. while at 10 Hz, Qseis is often as high as 1000 due to the
578 Conclusions

combined effect of deeper crustal stress levels, and a pre- 10.14


sumed more massive rock mass.
The more recent possibility to record seismic events using
down-hole seismometers, to depths of 100’s of meters and
10.11 even to 2 and 3 km, has made it possible at some locali-
ties, notably near the San Andreas Fault, to avoid con-
The late coda is assumed to contain a variety of take-off tamination by seismic and man-made noise, and to avoid
angles from the source. The frequency dependence is pre- much of the ‘site effect’ and it’s unwanted influence on
sented in the form Qc  Qo f n, with component n often recordings. Amplification in the low velocity near surface,
increasing from pre- to post-event values. Component n with increased scattering and intrinsic losses, can have
may be just below to close to 1.0. A greater density of had a strong effect on all surface recordings. Attenuation
scatterers after a seismic event has been suggested as the at shallow depths beneath the surface layers, appears to be
reason for greater frequency dependence after a major little influenced by rock type, with similar low values of
event, which also suggests that frictional and partial satu- Qp and Qs in the first 1 to 2 km down wells in widely dif-
ration losses will be involved when new discontinuity sur- ferent rock types like tertiary sediments and crystalline
faces are being developed, or extended. basement. Perhaps each of these dissimilar rocks are nev-
ertheless stiff enough to have a partially common tectonic
imprint of jointing and faulting from the same neigh-
bouring major fault.
10.12

Temporal variations of coda Qc have been reported in the 10.15


periods before and following significant earthquakes.
From 30% to 300% reduction of coda Q prior to six M Qp values of 20, 30 and 55 from analysis of depth inter-
6.0 to 8.0 earthquakes have been reported, with an vals 0–300 m, 300–940 m, and 570–940 m, and Qp
increase in coda Q in only one case, and the possibility of increasing to 110–170 from 1 to 3 km depth, are also fol-
no obvious relation between coda Qc and a single major lowing the likely pattern of deformation moduli, when
earthquake. Temporal variations have more recently been Q p values are expressed as if GPa. The actual addition of
interpreted as due to migrating hypocentres, with an scattering and intrinsic losses are of course not distin-
example Qc reduction from 200 to 100 interpreted as a guished in such a simple modulus model, but since the
migration during the early period of activity, to after the modulus estimate actually includes block size, inter-block
activity. In rock mechanics terms, this could approximate friction and effective stress-to-strength ratio, it is in some
to changing from a higher modulus, higher stressed way sampling important causes and components of both
zone, to one with lower modulus and lower stress, after classes of attenuation. Reduced fracture density with
the event. depth, reduced friction losses on increasingly stressed
fractures, and scattering from a reduced number of frac-
tures, are reported reasons for the increasing Qseis values
10.13 with depth. These are all recipies for increased deforma-
tion modulus too, and a common experience in deep, as
At the M 7.2 Hyogoken Nanbu earthquake near Kobe, opposed to shallow tunnels.
seismic Q c was shown to have reduced following the
earthquake. At respective frequencies of 1.5 and 2.0 Hz,
Qc reduced from 81 to 68, and from 91 to 78, perhaps 10.16
representing seismic sampling at increasing depth with
higher frequency. At respective frequencies of 3.0 and For cross-continent studies of seismic coda, the so-called
4.0 Hz Qc reduced from 132 to 107 and from 186 to Lg phase common on regional short-period seismograms
162. All the above resemble feasible deformation modulus is often used. It is followed by the main coda, which can
trends, but obviously progressing to depths beyond exist- also be used for determining the magnitude of regional
ing Emass data. events. The Lg coda at 1 Hz is commonly referred to as
Conclusions 579

Qo, and has been used to describe deep crustal attenua- inter-bedding is found to be quite attenuating, with good
tion across continents and in component plates, orogenic differentiation between shale, limestone and clay. A major,
belts and major sedimentary basins. Comparing South dipping fault zone in a North Sea reservoir, between 1 km
and North America on this scale, one sees lower Qo in the and 2 km depth, was interpreted from down-going P-
Andean Belt (250–450), and in the Basin and Range waves as having a seismic Qp value as low as 45. Analytical
province west of the Rocky Mountains (250–300). By modelling suggested a strong inequality between the high
comparison, there are broad regions of very high Qo span- dynamic shear compliance, and the more conventional
ning the central Brazilian Shield and Amazonian Basins lower normal compliance. Low pseudo-static shear stiffness
(700–1100), and in the eastern region in the USA below for fault-scale structures is also a necessary input to
the Great Lakes. numerical distinct element models, where realistic subsi-
dence modelling is to match steep subsidence bowls, in
situations where continuum modelling fails to match
10.17 such measurements.
Higher values of Qo reportedly reflect the length of time
since the last major tectonic activity. Low Qo regions are 10.20
typical of seismically active regions with higher upper
mantle temperatures, or the presence of deep hydrother- In Chapter 13, there are numerous further sets of seis-
mal fluids, and variable amounts of fluids in major faults. mic Q data from laboratory-scale testing, in which the
Younger sediments cause local reductions of Qo due to velocity and attenuation data are given side-by-side.
their contained fluids, while older sedimentary rocks Tests at reservoir confining stress levels and with numer-
which have lost fluid, are less attenuating. A shallow sed- ous fluids (brine, gas, oil) will also be found. The rock
iment model for explaining shallow and lateral variation physics data also extends into velocity and attenuation
of attenuation shows Qseis values of 30, 50, 75 and 100 testing of the effects of bedding, foliation and induced
for sandstones and shales from 0–100 m, 100–300 m, fracturing, and the anisotropy caused by these features.
300–600 m and 600 m. Units of GPa are again sug- In Chapter 15, the treatment of poro-elastic modelling
gested in broad terms. also addresses the dispersive and anisotropic nature of
seismic Q, as caused by the coupled dynamic behaviour
10.18 of hydraulically connected equant pores, microcracks
and aligned sets of fractures of various sizes
Seismic Q has become increasingly important in hydro-
carbon exploration due to the improved sensitivity to
degree of saturation of the ratio Qs/Qp compared to the Chapter 11 Velocity structure of the
velocity ratio Vp/Vs. The presence of low seismic Q in earth’s crust
petroleum reservoirs can reflect the presence of over-pres-
sure, of fracturing, and oil. Values of seismic Q as low as 11.1
10 to 40 at more than 4 km depth are good indications of
favourable properties, likewise are pay-zones with over- The uppermost 5 km of the crust shows a rapid increase in
pressured gas giving seismic Q as low as 10 to 40, when deformation modulus and density, as pore space and
the remainder of the sequence shows seismic Q between joints are closed. However, the thermal expansion partly
about 50 and 130. balances the increase in seismic velocity, and P-wave veloc-
ities above about 6.5 km/s do not appear to be common
here. The crust usually varies from 20 to 60 km in thick-
10.19 ness beneath continents, while the oceanic crust is much
thinner, and is usually about 6 to 7 km thick, beneath an
In reservoirs where cross-well tomography is performed, average water depth of 4.5 km. The typical crustal P-wave
there is a possibility of improved definition of the struc- velocity range is 6.0 to 6.8 km/s. In the upper 5 km, and
tures, due to the high frequency and multiple ray-path excluding sedimentary rock with Vp  5 km/s, the most
coverage. Transversely isotropic ‘layer-cake’ sedimentary typical range of velocity is 6.0 to 6.2 km/s.
580 Conclusions

11.2 on a scale larger than the laboratory samples, also giving


high permeability throughout the drilled section.
An almost linear velocity-depth gradient between 5 and
25 km for the average crust shows a gradient of about
0.6/20  0.03 s1, while the gradient between 5 and 11.6
10 km is approximately 0.5/5  0.1 s1. Velocity-depth
gradients are more than an order of magnitude steeper The P-wave velocity of the sub-ocean crust at and near
than this in the upper kilometre, and much steeper again ridge crests exhibit an increase in velocity with age.
in the upper 25–100 m. Gradients are even more pro- Numerous results from the Atlantic and Pacific mid-
nounced in the upper 25 m of weathering-effected rock ocean ridge studies show an obvious link between Vp and
mass. The reduced gradient at great depth is due to the age, with an increase of Vp of 2–3 km/s, in the first 40
expansion effect caused by increased temperature. million years. Deeper older layers do not show systematic
increase in velocity.

11.3
11.7
Continental-scale velocity-depth profiles show strong
visual resemblance to near-surface refraction seismic pro- Early laboratory testing of oceanic basement rocks from
files to 50 m depth in place of 50 km depth. A uniform deep drilling in the mid-Atlantic ridge highlighted the
subtraction of about 2 km/s velocity, and a scale reduction discrepancy between laboratory seismic properties and
of 1:1000 gives an almost indistinguishable result from in situ, bulk velocities obtained from seismic refraction.
the laterally-varying and depth-varying shallow refraction Confining pressures of 20, 50, and even 200 MPa
seismic obtained at a rock engineering site. were used in early on-board velocity measurements.
Laboratory velocities of 5.5 to 6.5 km/s were typically
obtained with 50 MPa confining pressure. The compa-
11.4 rable, shallow depth refraction-seismic inferred veloc-
ity-depth profiles showed only 2.5 to 3.5 km/s in situ
For many years, sub-oceanic marine seismic refraction velocities, an apparent discrepancy of about 3 km/s rel-
profiles were interpreted as a small number of layers ative to the intact rock, but in fact, some of this differ-
(Layer 2 to 2 km, Layer 3 to 7 km), separated by planar ence was due to the excessively high confining pressures
interfaces, with a constant velocity assumption for each applied to the intact samples.
layer. Homogeneous layering assumptions from the 1960s
were first replaced by much finer layering (2A, 2B, 2C and
3A and 3B), and then in the mid-1970s by continuous 11.8
gradients in velocity. The first geophysical downhole sonic
logging data for oceanic crustal material was near a portion It was subsequently realised that it must be the rapidly
of the mid-Atlantic ridge in a leg of the Deep Sea Drilling increasing (from zero) effective stresses, not the assumed
Project. Velocities were typically from 1.5 to 4.8 km/s in external ocean-depth loads, that were acting on the shal-
the upper 200 m of oceanic Layer 2A. Interpreted porosi- low sub-ocean crust, that was causing the rapid increase in
ties of 13 to 41% were unexpectedly high. velocity with depth. In other words, the velocity gradients
were similar to what is found at the earth’s surface, with
rock and (effective) fluid loads both increasing from zero.
11.5 The theory of effective stress was apparently late in being
adopted in this hostile sub-ocean environment.
The reasons for high porosities were interpreted as being
due to a combination of sediments, rubble, and solid
basalt in contrast to the compact nature of basalt sam- 11.9
ples used in laboratory tests, which often showed Vp
between 5.5 and 6 km/s and porosities from only about 2 Increasing effective stress from the ocean floor was respon-
to 8%. Open fractures and voids were assumed to exist sible for about 4 s1 velocity-depth gradients, together
Conclusions 581

with presumed 4 to 5% per 100 m porosity reduction, as interspersed by a majority of permeable and therefore
seen in the first 200 to 300 m of sub-sea layer 2A. The dif- low effective-stress-loaded permeable blocks.
ferences between in situ velocity measurements in the
shallow oceanic crust and the higher matrix velocities
measured at suitable (low) effective stress levels, was not 11.13
only caused by moderate changes to the matrix porosity,
but also by low aspect ratio jointing and fracturing, which The Vp-Q-value-porosity-depth model uses a plotting
was more stress-sensitive. format that can readily be compared with the oceanic
crust fracture zone data of Layer 2A and 2B. Strong
similarity with oceanic data is seen, probably because
11.10 the rock quality Q-value specifically represents the ‘soft
porosity’ or jointing, and models the effect on velocity
Very high velocity gradients, similar to the above, are of gradual joint closure with depth. The very steep
typically experienced at the earth’s surface because the Vp-depth gradients typically seen close to the ocean floor,
rock quality Q-value, as well as the effective stress, are in the first few hundreds of meters of the new crust, can
both increasing rapidly with depth. We may have a be analysed with this near-surface based empirical
near-surface Q  0.1 followed rapidly by Q  1 and method, which was developed from near-surface civil
then Q  10, suggesting nominal ‘near-surface’ theo- engineering projects. ‘Curve-jumping’ is needed to
retical increases in Vp from 2.5 to 3.5 to 4.5 km/s, with explain the supposedly ‘anomalously high’ gradients
the additional effect of increasing depth and therefore through Layers 2A and 2B. These can be modelled by
increasing effective stress. This dual effect requires ‘curve assuming increased Q-values for deeper, older material.
jumping’ in the Q-Vp-depth relation.
11.14
11.11 East Pacific rise studies were made by many of the
researchers known for their mid-Atlantic ridge studies. A
A direct measurement of Upper Oceanic Crust P-wave linear velocity-depth gradient in the upper 500 to 800 m
attenuation was described in 1990, using seafloor of young (0 to 4 m.y.) oceanic crust on the flanks of the
hydrophones and large explosive sources. The site was East Pacific Rise was initially assumed, and an average
on 0.4 m.y. old crust, and had a seafloor velocity of gradient of between 3.0 and 3.5 s1 for the upper 0.5 to
2.7 km/s, which increased uniformly to 5.6 km/s at 0.8 km of oceanic crust was estimated, with seabed veloc-
680 m depth. Gradients as high as 4.6 s1 near the sur- ities ranging from as little as 1.9 to 2.7 km/s. The evi-
face and 4.1 s1 at greater depth were estimated. Values dence of very low velocities in the upper-most oceanic
of seismic Qp varied from 4 to 275, but mostly clus- crust was consistent with visual and photographic evi-
tered between about 10 and 20 in the upper 100 m, dence from submersibles, of pervasive fracturing in mid-
similar again to expected deformation moduli in GPa. ocean ridge crustal regions, where the basalt layer was
exposed. Low velocities were also consistent with drilling
and logging results that showed high porosity.
11.12

A consistent increase of Qp with depth was not found, 11.15


but several sets of data for 1/Qp did show a ‘stepped’
trend of 1/Qp reducing with depth in the first 500 m. The problems posed by zero-age oceanic crust with
Values of Qp were mostly from about 8 to 50 in this Vp  2 km/s, compared to about 6 km/s for intact basalt
depth range. The sudden steps up, and down from, continued to provide challenges for theoreticians and
very high Qp values like 200–300, even negative 1/Qp practitioners working on the origin, formation and struc-
steps, leads one to question whether the early ship-board ture of mid-oceanic crust. Low aspect ratio cracks, and
triaxial test routines had an element of (local) correctness. their reduced frequency of occurrence and reduction in
Some volumes of intact basalt could perhaps be subject aperture with depth, and probable sealing with hydrother-
to high 30 MPa rock-plus-water confinement loads, mal minerals in the case of older oceanic crust, were some
582 Conclusions

of the variables confronting those researching the variable such very high seismic Q values appear here at ‘shallow’
structure of mid-ocean crusts. It was theorised that young (rock) depth compared to generally greater rock depths
120 ka material with Vp  2.5 km/s, must have a porosity on-land. With a thin, warm sub-ocean crust, permeabil-
of between 24 and 34%. Slower Vp  2.2 km/s zero-age ity may be compromised at relatively shallower depths
crust was theorised to have a porosity of between 26 than under tectonically deformed continental crust.
and 43%.

11.19
11.16
Age effect reviews for both mid-Atlantic Ridge and
At the beginning of the 1980s, a sub-ocean Deep Sea Pacific Rise data show that most of the age-dependent
Drilling Project borehole in the Costa Rica Ridge area, increase in seismic velocity occurs ‘rapidly’ with velocities
made it possible to correlate core, usually of low % recov- nearly doubling in less than 10 million years. Layer 2A
ery, with downhole sonic logs, borehole televiewer logs, appears to persist as a low velocity capping of the ocean
and permeability test results. This was first performed to crust, even when more than 15 m.y. old. The trend for
a depth of 1 km, through layers 2A, 2B and 2C. In the increased velocities as age increases is clearly shown by the
0–100 m, 100–650 m and 650–1000 m depth zones, P- statistics. A velocity plateau, averaging about 4.3 km/s is
wave velocities were a ‘familiar’ 3.7, 4.8 and 5.6 km/s, and indicated, beyond about 7 m.y. There is a clear link
permeabilities were likewise a ‘familiar’ 106–107, between hydrothermal alteration and seismic velocity
108–109 and 109–1010 m/s. Based on vertical bore- increase, due to deposition of minerals first in the
hole logging, which would be biased against vertical struc- thinnest cracks and joints. Hydrothermal void filling
ture, the upper 50 metres was found to contain numerous causes a simultaneous increase in velocity and reduction
horizontal to sub-horizontal fractures, thick basalt flow in hydraulic conductivity, therefore reducing heat flow to
units, and thin interbeds of pillow structures. the ocean floor.

11.17
11.20
Large scale three-dimensional tomography was performed
on the East Pacific Rise sub-ocean crust at the end of the In regions with significant sediment cover, the previously
80’s, at the location of a fast spreading ridge. This was open seawater convection cooling system is hindered, and
characterised by a sharp upper-crustal to mid-crustal temperatures rise, thereby accelerating the formation of
velocity inversion some 1.5 to 2 km below the seafloor. secondary minerals and porosity sealing. The on-axis, zero
This was presumed to be the roof of an axial magma lense, age upper crustal permeability has been deduced to be
with an assumed few percent of melt. Contours of seismic about 6  105 m/s, decreasing to about 7  107 m/s
Q at 4 km depth, showed values of Qseis of 25 and 33 within 6 m.y. Seismic velocities for crust of the same age
nearest the ‘magma’, and values of 50 and 100 at 2 to 3 km are about 2.2 km/s and 4.0 km/s. Permeability may reduce
off-ridge distances, in older crust. Values of Qseis as low as to about 107 m/s or less, by the time the crust is old
8 and 10 have been measured at the Costa Rica Ridge enough to have reached the approx. 4.3 km/s ‘plateau’.
area, caused by local faulting or similar features at 1⁄2 and
1 km depth in the much referred Hole 504B.
11.21

11.18 Parallels to the hydrothermal mineral filling of fractures,


and increased velocities may be gleaned from civil engi-
Inversions for individual receivers showed that seismic neering, where the sealing of jointed rock by pre-grouting
Q increased from average values of 40–50 in the upper with fine-grained micro-cements ahead of tunnels, or the
1 km, to at least 500–1000 at depth greater than 2 to use of industrial, coarser grain size cements in more per-
3 km. One may speculate that this abrupt increase could meable dam foundations, are common ways of both seal-
be due to an undrained increment of effective vertical ing and improving rock mass properties. Rotations and
stress of about 30 MPa representing the ocean load, since magnitude reductions of the three principal permeability
Conclusions 583

tensors, when conducting multiple-borehole 3D hydroto- pore pressure for conversion of the three principal stresses
mography before and after grouting have been docu- to effective stresses. The appropriate selection of wellbore
mented. ‘temporary support’ in the form of mud pressure, using
variable mud weight, determines the state of the bore-
hole wall in the different lithologies, prior to setting and
11.22 cementing the casing.

These tensor or principal value rotations are interpreted as


due to successive sealing of the most permeable and least 12.3
stressed joint sets. This process can also be interpreted by
small changes in five or six of the rock quality Q-value Due to various opinions about an alteration zone around
parameters (ratings in Appendix A), which may cause deeper wells, there is now widespread acceptance of the
some dramatic potential improvements in the rock mass need for logging while drilling (LWD) with monopole and
properties such as increased modulus of deformation, seis- dipole tools, to obtain ‘early’ velocity responses, which
mic velocity, and frictional and cohesive strength. There is may differ significantly from subsequent wireline logging.
a degree of correspondence to sub-ocean velocity increase, The differences are probably due to stress-fracturing,
and their expected geomechanical effects. increased permeability, and consequently accelerated
mud-filtrate invasion. A near-wellbore, tangentially-dis-
tributed log-spiral-type discontinuum, in case of an insuffi-
Chapter 12 Rock stress, pore ciently mud-weight supported weaker formation, may
pressure, borehole need to be considered when interpreting the two sets of
stability and sonic seismic data.
logging

12.1
12.4
Hydrocarbon-bearing rocks rely on pore-space and
permeability for the possibility of having recoverable It is important to consider the components and modifiers
reserves that can be produced at a well. The necessary of the most fundamental of reservoir parameters, namely
migration of the hydrocarbons from source rocks into the effective stress magnitude. The rock stress and its varia-
potential entrapment structures, without escape to the tions with direction, depth and location, and the pore
atmosphere, adds to the adverse statistics of hydrocarbon pressure and sometimes over-pressure which are influ-
discoveries. Too close to the surface the sealing properties enced by compaction and also by fluid type, are the major
of shale, salt or clay-smear in faults, may have been com- boundary conditions. Their relative magnitudes affect
promised by lack of plasticity and too high permeability. both the laboratory test simulations, the drilling pro-
Too deep, the pore space and permeability of the reser- gramme, the production planning, and the reservoir pro-
voir may be compromised, giving a reduced reserve and duction and depletion, possibly for 80 years or more in
the need for permeability enhancement and gradient a large reservoir.
enhancement, or a decision for non-development.

12.5
12.2
To prevent hydraulic fracturing by high mud-weights,
Besides reservoir access for production testing, a drill- which are needed where there is overpressure, casing will
hole is used for sonic logging and selected side-wall and be set to protect the overlying units from fracturing. A
regular core recovery, to better define the properties of the change from a pressure-depth axis, to mud-weight-depth
different lithologies, seals and reservoir rocks. Rock reacts format is preferred by drillers, who speak of mud-weight
to the drilling of boreholes with a complex interaction of in lb/gal, and try to steer this between the pore pressure
rock stress and strength magnitudes, plus the anisotropies gradient and the fracture-gradient. Resistivity, velocity,
of each, and is affected by the necessary subtraction of and density depth-trends will each suffer various degrees
584 Conclusions

of deviation from the norm, when there is over-pressure minimum of the three principal rock stress magnitudes in
that changes the effective stress. the shale or salt will therefore often exceed the minimum
stress in the reservoir sandstone by up to several MPa. This
is desirable for hydrocarbon containment, and also for
12.6 vertically limiting massive hydraulic fracture treatments.
Rocks such as granite, limestones, stronger sandstones
The fluid pressure at a well is the sum of normal hydro- and stronger shales (which are thereby poor seals), toler-
static pressure, plus over-pressure, plus a buoyancy effect ate differential stress much better than weaker shales and
caused by the reduced density of any petroleum that is almost plastic salt rocks.
present. Since over-pressure and the presence of petroleum
products both increase the pore pressure, the effective
12.10
stress will also be reduced, which will have the effect of
causing a reduction in velocity.
Shale or salt may, if encountered at sufficient depth during
drilling, require the support of an active mud-weight to
prevent creep or squeezing. The drillers choice of mud-
12.7
weight, or the setting of protective casing, becomes critical
where support of the well is needed adjacent to a reservoir
Over-pressure commonly occurs where low permeability
rock like sandstone or fractured limestone or chalk, which
layers such as shale prevent fluid from escaping as rapidly
would tend to have a minimum rock stress less than that of
as pore space compacts. Excess pressure in relation to
these weaker, sealing ‘plastic’ layers. The reservoir horizons
hydrostatic then builds up as newly deposited sediments
could potentially fracture, or have a permeable joint under
cause squeezing of the trapped pore fluids, which could
lower normal stress than the mud-weight needed to keep
be water, oil or gas or two or three of these close together.
the plastic materials from squeezing and jamming the
Models for basin-evolution show that pore pressure
drill-string. Invasion of mud or lost circulation, into any
effects are seismically visible when the effective pressure is
reservoir horizon is obviously very undesirable.
typically less than about 15 MPa, A small % conversion of
live oil to gas is sufficient to make the pore pressure equal
to the confining pressure. The large changes of predicted 12.11
velocity are caused by the fact that the dry rock moduli
are strongly affected by low effective pressures. As one approaches the surface, inter-bedded rock types
resembling reservoir sequences, show the reverse of
the previously discussed differential stress intolerance,
12.8 because the weaker rocks are no longer ‘over-stressed’.
Furthermore, because of their lower deformation moduli,
Minimum rock stress estimation by mini-hydrofracing, if they attract lower stresses from a given horizontal stress
not possible in an open-hole situation, is done by seating field. Hydraulic fracturing tests therefore may give indica-
the double-packers on either side of shaped-charge perfo- tions of low Ko ratios ( h min/ v) in the weaker materials
rations of the casing. This is done in the reservoir intervals, like shale and siltstone, and higher values in sandstones
and also in the cap-rock interval, to determine the mini- and limestones.
mum stress difference. Interbedded ‘brittle’ layers like
sandstone, and ‘plastic’ layers like shale or salt, will usually
exhibit different minimum principal stress levels. This 12.12
may be an additional reason for oscillating sonic log
records in such interbedded strata. This reversal of Ko trends at a certain depth (it was meas-
ured from 100–150 m depth, but might apply from 0 to
500 m), may have implications when comparing stress-
12.9 induced velocity anisotropy and sonic log velocity ‘oscil-
lation’ near-surface and at greater depth. This reversed
Shale and salt-rocks may have insufficient shear strength behaviour also needs to be considered when evaluating
to tolerate a significant principal stress difference. The the applicability of shallow borehole seismic testing to
Conclusions 585

reservoir holes perhaps an order of magnitude deeper, temporary support costs by under-supporting, using for
with corresponding reversed Ko behaviour. There may example a constant single layer of sprayed concrete, where
also be consequences for the relative magnitudes of two layers were actually needed locally.
attenuation, as both lower Ko and lower stress levels near
the surface, will tend to enhance attenuation. Thus Qp
and Qs values must be expected to be lower, and exhibit 12.16
more anisotropy near-surface than at depth.
Due to the smaller size of wells and the use of mud for
hole support, the recognition of the behavioural data that
12.13 can be extracted from anisotropic stress effects on small-
scale EDZ round wells, may possibly not be used in the
The hydrocarbon reservoir exploration and production petroleum industry, to the extent it can be used in tunnel
industry has long been aware that borehole deformation engineering. Borehole ellipticity, a much-used historical
and failure modes are an important ‘complication’ con- indicator of the minimum horizontal stress axis, is the
cerning the interpretation of sonic-logging of wells. surface expression of effects behind the ellipticity. In tun-
There are now acoustic dipole and monopole shear- nels it is easy to see the effects of structure-induced
wave producing logging devices that can be used in a wedge release, or stress-fractured ‘lenses’ of rock. It is also
logging while drilling LWD mode, that acquire responses possible to install multiple-position borehole extensome-
from more than one hundred wave forms, some tens of ters (MPBX), in tunnels and in vertical shafts, to meas-
meters behind the drill-bit, in order to delineate forma- ure the anisotropic radial-distribution of deformation,
tion fracturing response, and virgin conditions further thereby giving deformation moduli as a function of
from the walls of the wells, before additional ‘alteration’ direction. Velocity variations and permeability variations
has occurred from stress and/or mud-filtrate invasion, as a function of position and radial depth around a tun-
as often seen in subsequent wireline logging, when the nel or shaft, can also be determined, thereby relating
drill-string is removed. these parameters to eventual stress anisotropy.

12.14 12.17

There are possibilities for local velocity (and seismic Q) There is a strong likelihood that mini-EDZ in the weaker,
enhancement due to tangential (and diametrically-oppo- less well mud-supported zones, have reduced, radial-
site) stress increase in the case of competent rock like dependent velocity, due to failure and deformation in the
limestones, or low porosity sandstones. In the case of over-stressed zones. Stronger inter-beds could show an
over-stressed, fractured (‘dog-eared’) sections of rock, and opposite trend due to tangential stress enhancement of
especially in the case of incompetent rocks like shales, the velocities. Mini-EDZ that might penetrate several
reduction of velocity (and of seismic Q) will occur locally, diameters can be detected, and circumvented by deeper
due to the mini-EDZ (excavation disturbed zone) that sensing, shear-wave based, dipole logging tools. The prob-
form as a result of drilling and possible over-stressing. able discontinuum caused by log-spiral shearing is often
referred to only as ‘shale alteration’. Fabric and jointing
and bedding planes, may also affect the progress rates for
12.15 mud-filtrate invasion. The geomechanics of borehole
deformation and over-stress, and its coupled MHT effect
The mini-EDZ may mean the development of a log-spi- on permeability, mud-filtration, and LWD-to-wireline
rally sheared discontinuum, based on physical and numer- logging differences, can be quantified in approximate
ical modelling results, where the starting point was a terms.
continuum. Stress reduction in the radial direction, which
may be azimuthally varying, will tend to locally reduce the
velocity, and thereby also the seismic Q. It is impossible to 12.19
support each lithology with the ideal mud-weight, so
some suffer the consequences, just as occurs in weak zones Deeper penetration of mini-EDZ, representing ‘shale
in a tunnel where the contractor might be trying to reduce alteration’, may be the reason for a serious potential
586 Conclusions

contrast in logging results, when comparing the 1–2 followed by frictional mobilization at larger strain. In
weeks later result of wireline logging, with the few modified Mohr-Coulomb terms it is a case of ‘c then tan
hours delay represented by LWD, or logging while ’, not ‘c plus tan ’. Numerical models that are pro-
drilling. More recent shear-wave anisotropy based log- grammed, or manually-steered, to dissipate cohesion
ging, is capable of imaging a volume of up to several while mobilizing friction, are capable of matching physi-
borehole diameters away from the wall, therefore cally observed behaviour. Non-linear fracture mechanics
beyond the stress-related fracturing and mud-filtrate boundary-element based modelling seems to mirror real-
invasion or ‘shale alteration’, thereby giving presumed ity extremely well, with log-spiral type fracture develop-
‘virgin’ formation attributes as well. ment that dissipates over-stress.

12.20 12.23

Variable azimuth drilling in test blocks under 3D stress Different degrees of ‘log-spiral-type’ shear failure are
states, gives failure modes that cannot be obtained when demonstrated, depending on the ‘disturbance’ to the
loading a test block with a pre-drilled hole. Deep log-spi- isotropic stress distributions, caused by different amounts
ral shear failure surfaces have been demonstrated in weak of jointing or fissuring close to the hole. These geologi-
cemented-sand blocks, when the major principal stress cal features dissipate some of the highest, near-wall tan-
was about eight to ten times higher than the uniaxial gential stresses seen in elastic isotropic analyses. When
strength, with the minor and intermediate principal modelling medium strong brittle rock, principal
stresses of 60% or 80% of the maximum. This level of stresses of only about 35–40% of the uniaxial strengths
over-stress is easily reached in deep wells in relation to are needed to start fracturing in the form of initial ‘dog-
shale and salt rocks. earing’. A brittle sandstone of 50 MPa UCS would be
acted on by an equivalent ratio of strength to stress
beyond about 1200–1300 m depth, with standard den-
12.21 sity and pore pressure assumptions, considering a
H max value no larger than the vertical effective stress.
With extreme weakness, the failure mode may be non- The existence of dog-earing may provide more informa-
dilatant, and actually contracting-with-shear, or even flow, tion about formation properties than the simple registra-
in the case of clay-rich materials. There are distinctive dif- tion of stress-induced ellipticity in a four-arm calliper log.
ferences between stress-induced failure of hard dilatant
brittle rocks, giving extensional splitting and subsequent
crushing or comminution of the rock in the sharp ends of 12.24
diametrically opposite ‘V-shaped’ corners. In the case of
failure in intermediate strength and less dilatant rocks, the Due to the influence of deformation of ‘soft’ as opposed to
traditional ‘dog-earing’ takes on a different shape resem- ‘hard’ porosity, a borehole for hydrocarbon exploration
bling localized log-spiral shear failure surfaces. Each of that penetrates variably jointed and faulted ground, will
these modes can be demonstrated in physical simulations, actually experience variable small amounts of deformation,
and with suitable choice of numerical model, though not due to different degrees of joint closure, joint opening,
with conventional rock failure criteria. and joint shearing. There will also be the pseudo-elastic
response, due to both loading (at the diametrically-oppo-
site max locations) and potential unloading (at the dia-
12.22 metrically-opposite min locations) of the matrix as well
as the joints, the latter usually dissipating some of the the-
The actual modes of physical behaviour experienced by oretical (isotropic, elastic) peaks of maximum and mini-
boreholes and tunnels, are unlikely to be predictable mum tangential stress. This process will occur even with a
when modelling with conventional Mohr-Coulomb type constant mud-weight, since the mud – unlike rock bolts
(c  tan ) shear strength criteria, because intact rock in a tunnel – cannot prevent joint movements of unequal
tends to fail first by loss of continuity at small strain, magnitude at different points around the opening,
caused by loss of local tensile or cohesive strength, although the mud may help to make them very small.
Conclusions 587

12.25 by a reduced elastic modulus in an annular zone around


the borehole, particularly in soft formations such as shales
When assessing the rock quality of the walls and arches and shaly sands. Entrapment of wave-fronts in the lower
of tunnels, the observed rock, which is the visible part of modulus damage zones results in ‘bi-compressional
the tunnel-scale EDZ, is classified (using the Q-system: arrivals’, or second arrival compressional waves. The bi-
see Appendix A), in order to select appropriate rock rein- compressional arrival is a phantom arrival too fast to be a
forcement (grouted rock bolts) and tunnel support shear wave, and actually caused by trapping of the wave-
(sprayed, steel-fibre-reinforced concrete). The latter is front by the low-modulus damage zone.
the equivalent of the borehole mud pressure, and is badly
needed in a complete load-bearing ring, in the rapidly
deteriorating and deforming clay-bearing zones, in 12.29
order to control deformation and prevent local tunnel
collapse. Outside the tunnel EDZ, the rock mass would Orthogonal dipole transmitters, and the multiple receiver
be characterized as a better quality rock mass. pairs, which are aligned in orthogonal directions, measure
the components of slowness in any direction within planes
perpendicular to the borehole. They use the principal of
12.26 shear-wave splitting, and polarization, one of the most
valuable of all seismic anisotropy properties for fracture
If one performs both sonic logging and azimuthal dipole and fracture-fluid investigations. The slow direction (as
sonic logging in a well, the borehole mini-EDZ can be also with P-waves) is perpendicular to the fracturing –
classified by the one tool, while the hydrocarbon-bearing which could be microcracks, cracks, joints or faults –
or reservoir sealing formation away from the immediate according to the scale of example considered. The rotated
influence of the hole, can be characterized by the other direction of the fastest shear waves becomes the fast-shear
tool, and show higher velocity and higher seismic Q as tool azimuth. Both the acoustic time anisotropy and the
well. It appears that this familiar rock mechanics EDZ slowness anisotropy are sensitive to properties deeper
logic is effectively being applied in modern well logging, within the formation than the superficial effects caused by
with its multi-wave-form acquisition. drilling.

12.27 12.30

Dipole transmitter tools are designed to generate flex- The fact that shear wave anisotropy allows the investiga-
ural waves. Flexural waves are shear waves that are tion of a volume of the formation up to several diameters
polarized into fast and slow directions, and penetrate from the borehole axis, means that it can sense jointing,
several hole-diameters into the formation, thereby reveal- and stress-induced fracturing, that are missed by conven-
ing potential stress-induced ‘alteration’, and/or drilling tional logging tools. This means that it is particularly use-
mud-induced alteration. The need for these tools con- ful for registering the additional jointing and fracturing
firms many of the foregoing suspicions that what we that tends to be present on either side of a fault.
have termed mini-EDZ, are indeed a source of concern
in certain formations, and more importantly, that these
‘alteration zones’ can be detected and seismically classi- 12.31
fied. The shear-wave analysis can also be used when
characterizing the formation beyond the damage zone. LWD with dipole shear-wave anisotropy analysis is
available almost in real-time, some hours behind the
drill-bit. It has proved very useful when applied to
12.28 drilling of horizontal well sections, designed to intersect
a maximum amount of structure. Simultaneously one
The ‘altered zone’ around the borehole may continue to can avoid the less favourable parallel to H max hole
develop during the week or so that may separate the two direction. Early warning is also given while drilling in
types of logging. The later wireline log may be influenced formations with rapidly changing pore pressure. LWD
588 Conclusions

is then an important aid in choosing appropriate mud 12.35


pressures. The use of wireline dipole logging in vertical
holes, and pipe-conveyed dipole tools for deviated and Mud cooling systems can be applied, which help to
horizontal wells, has given reservoir geophysicists reduce the likelihood of compressive stress-induced fail-
improved means of calibrating the responses of their ure, as matrix contraction reduces the maximum tangen-
rock physics based reservoir models, against these tial stress concentration. On the adverse side, there is an
small-scale, but in situ measurements. increasing possibility of tensile failure and mud loss. So-
called wellbore ‘ballooning’ represents such mud loss,
but some of the mud is returned by subsequent heating.
12.32 If there is no risk of compressive stress failure, then high
mud temperatures can be used to control the hydraulic
Synthetic seismograms often do not correlate with fracturing gradient. Shale and salt rocks pose particular
measured seismograms, when correlating seismic data challenges in HPHT environments.
with acoustic logs. Formation properties inferred from
wireline logging measurements may not reflect the true
properties, so a realistic description of the mud-filtrate Chapter 13 Rock physics at
invaded damage zone is important for processing and laboratory scale
interpretation of the logs. A problem is caused by the
invaded or ‘altered’ zone being deeper than that illumi- 13.1
nated by the logging tool, meaning that the velocities
will not reflect those of the formation, but of the dam- The major exploration-related goals of rock physics
aged zone, therefore requiring corrections. research have recently been summed up by King as: ‘to
understand how lithology, porosity, confining stress and
12.33 pore pressure, pore fluid type and saturation, anisotropy
and degree of fracturing, temperature, and frequency
A standard approach is to correct the acoustic logs via a influence the velocities and attenuation of compressional
Biot-Gassmann fluid substitution, to free sonic logs from P- and S-waves in sedimentary rocks’. At the end of this
mud-filtrate invasion effects. It is assumed that the list ‘and vice versa’ was added, emphasising the interac-
measured velocities are those of the invaded zone, satu- tive and complex nature of the reality. The remaining
rated with mud filtrate. By displacing the saturation challenges were summed up as ‘relationships between
fluid theoretically, new velocities are obtained, and taken attenuation, anisotropy, fractures and fluid flow, and
as the virgin formation velocities. Despite theoretical determining these relationships across the frequency spec-
removal of the saturating fluid, one possibly should note trum of core, log and seismic measurements’.
from Chapter 15 the failure of the porous-medium
based Biot-Gassmann method in the case of split shear-
wave data in fractured as opposed to unfractured porous 13.2
formations.
It was recognized long ago that seismic velocities of
rocks were strongly influenced by micro-cracks, and that
12.34 seismic attributes representative of the intrinsic mineral-
ogy and porosity, could only be obtained by applying
Ultra high pressure and high temperature (HPHT) wells pressure to the rocks. Much of the rock physics under-
that are increasingly relevant with high petroleum prices standing of reservoir rock behaviour, of both matrix
have several unfavourable effects on well stability. One is and joints or fractures, has therefore to be achieved at
that the rock skeleton may bear a proportionally greater elevated pressure. The importance of elevated tempera-
load, as the effective stress parameter () becomes less ture is also well recognised, but is less frequently an
than 1.0 (as per Terzaghi), when stress levels have caused experimental variable than it actually should be, espe-
reduced porosity and permeability. This higher effective cially if jointing is also sampled, as it should be in the
stress causes slower drilling rates. case of fractured reservoirs.
Conclusions 589

13.3 13.6

Age-depth relationships derived from well analysis in Theoretical modelling of porous rock behaviour has
sandstone-shale units have a certain grouping of veloci- been used on many occasions for examining the numer-
ties with age, due to variations of porosity and the ous factors affecting seismic velocities. Theoretical for-
resulting densities. Hard porosity in the form of pores mulations by Toksöz were used to represent the solid
tends to decrease with age and depth, while soft poros- matrix, and the assumed spherical to oblate pores, using
ity in the form of joints tends to increase with age due to widely varying aspect ratios to match numerous labora-
tectonic influences, but reduces strongly with depth. tory data. Small aspect ratios, or flatter voids, caused the
Only the hard porosity has a significant effect on density. greatest reductions to elastic moduli and velocities. The
A relatively ordered density-Vp trend for chalk, lime- properties of the saturating fluid (gas, oil or water) were
stone and dolomite is often seen, and reflects the simple found to produce greater effects on the compressional
mineralogy. Frequently widely scattered density-Vp data velocities than on the shear velocities. The P-wave veloci-
for sandstones is evidence of the variable mineralogy of ties were predicted to be higher, and of course were meas-
‘sandstones’, with 10–15% variation in density possible ured as higher, when the rock was saturated with water,
for the same velocity, particularly in the case of tight gas than when dry or gas-saturated.
sandstones. In contrast to these variations, Vp-Vs trends
are consistently uniform, as befits characterisation by
seismic waves. 13.7

When such theoretical models are fitted to P- and S-wave


velocities that have been measured at different pressures,
13.4
they are found to require pore shapes ranging from
spheres to very fine cracks (aspect ratios from 1 to 105)
An early compilation by Faust of well survey results from
for sandstones, limestones and granites, both under dry
some 500 petroleum wells in the USA and Canada,
and saturated states. As igneous rocks have low porosities,
included data from about 300 kilometres of well sections.
the pore shape has great influence on the elastic and seis-
The great majority of data was for mixed shale/sandstone/
mic properties, and dry and water-saturated behaviours
shale sections. A non-systematic comparison of shale and
are often very different as a result. Compressional veloci-
sand (sandstone) velocities revealed an average discrep-
ties are predicted and measured, as highest with brine sat-
ancy of only 350 ft/sec, or 106.7 m/s in velocity between
uration, and lowest with gas saturation. These differences
these two, frequently inter-bedded units, the sandstone
decline with increasing effective pressure. Poisson’s ratios
having the highest velocity by this small average margin.
for gas saturated rocks are predicted to be lower than for
Remarkably close Vp versus Vs trends for water-saturated
those with brine saturation, and this difference persists to
sandstones and shales, emphasise the remarkably similar
high pressures.
seismic velocity signatures of these two ‘dissimilar’ litholo-
gies, when in a compacted state. The necessity of using
attenuation, and other techniques, for seismically distin-
13.8
guishing these two most essential reservoir ‘partners’ is
clear.
From extensive data sets and modelling we can summa-
rize from Toksöz that Vp is likely to be lower: if low
water saturation, if dry or gas saturated (in the case of
13.5 flatter pores), if in the presence of some immiscible gas
(in brine), if higher porosity, if over-pressured, if at shal-
A useful means of separating such formations is to plot low depth, if containing thin pores, following several
impedance, or the product of velocity and density, versus cycles of freezing, if at room temperature, if at extremely
Vp/Vs. Variation of porosity within such a diagram is a high temperature. Vp is likely to be higher: with water
further means of distinguishing different formations, saturation, when dry or gas saturated (in the case of
once they have been identified. rounder pores), when saturated with brine, when there is
590 Conclusions

no immiscible gas, when of lower porosity, when under- greater drop in velocity (e.g. 5 km/s to 3 km/s) as clay con-
pressured, when at greater depth, when consisting of tent rises to 30% or more in a simultaneously increasingly
rounded pores, when frozen, when at low or moderately porous sandstone. The relative effects of clay content at
high temperature. frequencies of 10 Hz to 1 kHz (as used in seismic explo-
ration) or frequencies of 10 to 20 kHz (as used in borehole
logging), when imaging in situ sandstones, is less clear, due
to potential effects of anisotropy. The complex nature of
13.9 permeability, which depends on porosity, pore size distri-
bution, inter-connectedness of the pores, and tortuosity,
Water-flooding, and four dimensional seismic monitor- means that permeability may be severely compromised by
ing of its effects, relies on the strong dependence of increasing clay content. Variable clay content also typi-
velocity on temperature, and on the significant influence cally occurs in bedding-parallel layers, making for strongly
of the relative hydrocarbon and brine saturations. Both anisotropic permeabilities.
the matrix, but especially the joints, will also sense the
reduced effective stress, caused by the injection pressure
and by the matrix shrinkage caused by cooling. 13.13

When relating rock physics laboratory data to the in situ


13.10 reality of frequently inhomogeneous deposition cycles,
the technique of ‘fining-up’, to benefit from the more
Concerning the matrix behaviour in the case of oil sands, uniform sedimentary environment and diagenetic nature
the greatest effect of temperature is seen when there is of smaller deposition cycles, gives successively higher Vp-
100% oil saturation, with P-wave velocity reducing by as n% correlation coefficients. Sorting data into common
much as 40% (e.g. 3.4 to 2.0 km/s) from 20° to 150° C. sediment-compaction and cementation-history cate-
Least and almost zero effect is seen when there is either gories, using stratigraphy and other matching tech-
100% gas saturation or 100% brine saturation. As oil is niques, also has advantages when establishing relevant
removed from these sands, the velocities successively relationships between permeability and porosity, and
become independent of temperature, with roughly half between velocity and permeability. The log of permeabil-
the effect when 50% oil remains. In viscous tar sands ity may be linearly related to porosity (e.g. 30%: 1 darcy,
velocity may reduce by more than 50% (e.g. 2.4 to 20%: 10 md, 10%: 0.1 md).
1.1 km/s) when steam flooding reduces viscosity due to a
similar temperature rise.
13.14

Relationships between velocity and permeability, can be


13.11 developed more easily by making a more relevant
match of porosity and permeability with common sedi-
At low temperature the higher viscosity means that the ment compaction and cementation history. Without
oil cannot flow easily, so the dynamic measurement is doing this, velocity and permeability may show poor
on the high-frequency, high velocity, unrelaxed side of correlation. Simple ‘classification’ terms can be used.
the local-flow mechanism. As temperature increases, RQI is known as the reservoir quality index, and is com-
viscosity reduces, so fluid flows more easily, and veloc- posed of permeability (k) in units of millidarcies, and
ity therefore decreases since measurement is on the () the fractional porosity. The void ratio (), the ratio
relaxed side of the absorption/dispersion mechanism. of pore volume to solid volume, given by /(1  ),
links RQI and FZI, which is known as the flow zone
indicator. Rocks with FZI values within a narrow range
13.12 are found to belong to one hydraulic unit and have sim-
ilar flow properties. A semi-log plot of porosity versus
Ultrasonic measurement of the effect of clay content on log permeability tends to show similar FZI values plot-
the P-wave velocity of sandstones may show a 40% or ting together with distinct log K-VP trends. The FZI
Conclusions 591

classification can be extended to seismic parameters, 13.18


to obtain a stronger correlation between velocity and
permeability. The sensitive response of () close to the hydraulic frac-
turing limit may be the result of sealed, over-pressured
and under-compacted beds. There may also be additional
13.15 over-pressure due to increased oil-to-gas conversion at
zero effective stress. Vs is (locally) zero as the rock mass is
A uniform or smooth variation of velocity with degree of hydraulically fractured and load is born by the fluid.
saturation is strictly a function of an assumed or actual However Vp is not zero, therefore  approaches 0.5.
homogeneous distribution of saturation, which is possi-
ble with lithological uniformity. The more complex and
common effects of mixed lithological units, may create a 13.19
heterogeneous or patchy saturation distribution. This
gives different signatures during imbibition and It is well documented that pseudo-static moduli of
drainage. The drainage process creates a more heteroge- deformation can be significantly lower than the dynamic
neous distribution of saturation. Local full saturation of moduli predicted from P- and S-wave velocities.
the crack-like regions of the pore space tend to stiffen Knowledge of the non-linear elastic properties that are
these regions in relation to high frequency, but at low fre- largely responsible for the differences between dynamic
quency these ‘patches’ can drain to the less saturated pore and static moduli is essential for optimal drilling, effec-
space. The pore fluid lying in thin, compliant pores can tive well completions and efficient reservoir manage-
flow freely into the dry pore space, in a squirt-flow type ment. The frequency of measurement is all important for
of attenuation response. It does not therefore allow rein- the geophysical estimate obtained, since Eultrasonic 
forcement of the compliant part of the rock at low fre- Elog  Elow freq.  Estatic.
quencies, so velocities are low.

13.20
13.16
The development of attenuation as a means of improved
At high frequencies, pressure equilibrium cannot occur characterization of reservoir rocks is due to the disper-
because the pore fluid relaxation time is greater than the sive, frequency-dependent nature of seismic Q, and the
seismic wave period. Pore fluid in thin compliant pores is greater sensitivity of the ratio of Qs /Qp to fluid and par-
then effectively trapped, and it therefore reinforces the tial saturation than Vp/Vs. The expected reduction in Vp
otherwise compliant pore spaces, resulting in higher by reduced brine saturation and increased gas saturation
apparent modulus and velocity. in sandstones is matched initially, by greater attenuation
with Qp reducing from e.g. 30 to 10. At the far end of
the saturation scale, when samples become ‘room dry’ or
13.17 reach 100% saturation with nitrogen, the attenuation
reduces sharply, and Qp may reach a value of 50. This
Poisson’s ratios are anomalously high in cases of over-pres- is related to the eventual absence of squirt flow with
sure, where effective stress can approach the fracturing increased dryness.
(negative) side of the usual lithostatic and pore pressure
gradients. The aspect ratio of the cracks and pores and the
nature of the saturating fluid determine the magnitude of 13.21
(). Rocks containing mainly stiff, equi-dimensional
pores do not show major variations of () with effective Ultrasonic (0.1 to 1.0 MHz) laboratory tests on dry
stress. In saturated rocks the compliant pores become and water- or brine- saturated Berea sandstone of 16%
stiffened in relation to high frequency waves, so () porosity, have been used on numerous occasions for
changes less as effective stress increases. However, at low investigating how velocity and attenuation vary with
effective stress, when pore pressures are very high, the ‘differential’ pressure (confining minus pore pressure).
effective stress sensitivity is marked, and () increases. The P-wave velocities of this sandstone, when dry and
592 Conclusions

when brine-saturated, usually rise rapidly over the first 10 mechanisms. The saturated rocks always show much
to 20 MPa, eventually reaching a plateau with little stress stronger attenuation (lower Qp ) than the dry samples.
sensitivity. At high differential pressure such as 50 MPa, The effective stress level has greatest influence on attenua-
equivalent to reservoir conditions, there may be little dif- tion when samples are brine-saturated and at the lowest
ference between the dry and brine-saturated values of Vp. levels of effective stress. The behavioural contrast of Qp
By comparison Qp and Qs rise more consistently, even up compared to Vp, with the former’s greater sensitivity to
to high stress levels. They show less attenuation in the dry effective stress level, and to frequency, confirms the
or methane saturated states, than in the case of brine sat- importance of attenuation as an excellent diagnostic at
uration, where squirt losses can occur. reservoir stress levels.

13.22 13.25

Qs is often larger than Qp in the case of the dry and Differentiation of Qp values, has also been suggested as a
methane-saturated sandstone. In the case of brine satura- way of distinguishing gas and condensate from oil and
tion, there is a consistently wide separation of Qp and Qs water in sandstone reservoirs. In perfectly dry rocks, Qp is
(Qp  Qs). This represents an important means of track- very high. In fully liquid saturated rocks Qp is at an inter-
ing water-gas fronts during production. The increase of mediate level. In partially saturated rocks Qp is low.
Qp from e.g. 10 to 20 at low effective stress, to values as Magnitudes of Qp for sandstone reservoirs, based on well
high as 60 to 80 to 100 at effective pressures of 50 MPa log (i.e. sonic frequencies), are usually in the following
bears a strong resemblance to deformation modulus ranges: gas and gas-condensate bearing sandstone
increases. A lower range of increase seen with brine satu- 5  Qp  30, oil bearing sandstones 8  Qp  100,
ration, might be due to the weakening effect of brine. water bearing sandstones 9  Qp  100. The low Qp
values may be caused by low effective stress in the case of
over-pressure (or by fracturing). These ranges are remark-
13.23 ably similar to laboratory ultrasonic data, when a range of
effective stresses are applied.
Tests with higher porosity sandstones (range 20–25%)
tested in dry or brine-saturated states, in triaxial compres-
sion, over an effective stress range from 2.5 and 40 MPa, 13.26
using a range of frequencies (400 to 2000 kHz), show
important additional trends when the effect of frequency Classic Biot theory that accounts well for attenuation
is shown together with the effect of stress level. The in clay-free sandstones, fails by an order of magnitude
greater sensitivity of Qp than Vp as effective stress levels to account for the attenuation effect of clay content.
rise is seen as before. However the effect of increasing fre- Strong clay-related attenuation is assumed to be due to
quency shows negative velocity dispersion for the dry sam- viscous interaction between the clay particles and the
ples, meaning velocity decreasing with increasing pore fluid. Permeabilities are also strongly dependent on
frequency, while in marked contrast, the attenuation clay-content. The measurement of attenuation of com-
increases (Qp reduces), as the third to fourth power of fre- pressional waves in sandstones under confining pressures
quency. This is assumed to be evidence of scattering of 40 MPa at ultrasonic frequencies (0.5–1.5 MHz,
within the pore spaces between the grains. shows that intra-pore clay content is important in caus-
ing attenuation, and in modifying the permeability. Qp
may be as low as 10 with clay contents 10%, rising
13.24 up to several hundred when clay content is 1%.

Brine-saturated sandstones show slight, positive velocity


dispersion at the lower confining pressures, while attenua- 13.27
tion increases (Qp reduces), with only the first or second
power of frequency. This change in attenuation-frequency Systematic reduction of Qp from about 80–100 to about
dependence is taken as evidence of local fluid-flow loss 10–20 is seen as the percentage of compliant minerals in
Conclusions 593

sandstones and siltstones increases from a few percent to as low as 4 to 5, and 3 to 6, at near-surface sites, using
nearly 80%. This is attributed to ‘clay squirt flow’. This seismic frequencies. There is potential for strong squirt
mechanism may also be important at both seismic and flow attenuation with passage of low frequency seismic
sonic frequencies, in the case of larger scale geologic fea- waves, in dual porosity systems such as jointed limestones.
tures such as inter-bedded permeable and impermeable Smaller magnitudes of Qseis may be expected, due to the
layers. contrasting moduli at the different scales, with reduced
differences at higher stress levels.
13.28
13.31
Different rock types such as siltstones, sandstones and
limestones may also show a significant range of instanta- A degree of correlation is noticed between seismic Q and
neous sample deformation as a result of applying high the ‘static’ modulus of deformation, expressed in GPa
confining pressures. Their deformation moduli are clearly and readily estimated from rock quality Q. This modu-
different. Differences between rock types will be accentu- lus is stress- or depth-dependent, and may range from
ated when bedding and jointing is also present, causing about 1 to 150 GPa in the upper 1 kilometre, but most
increased attenuation, and greater sensitivity to effective frequently from 5 to 100 GPa. The components of the
stress. This sensitivity may even apply to velocities, which rock quality Q-value reflect many potential attenua-
tend to show much less sensitivity to confining pressure tion-causing factors, e.g. RQD/Jn for scattering due to
than seismic Q when samples are without jointing. relative block size, Jr/Ja concerning the frictional and con-
ductive properties of the joints that are expected to be rel-
13.29 evant for squirt flow, including loss mechanisms in clay,
Jw as a direct link to permeability, and SRF related to
Dual porosity limestone specimens, with micro-pores increased attenuation where stress is low, and reduced
and inter-particle macro-pores, show a weak trend for attenuation where stress is high.
higher attenuation and lower seismic Q when perme-
ability and total porosity are also larger. For example a
Qp of 10 roughly correlates with 1–10 mD, while a Qp 13.32
of 100 roughly correlates with 0.01–0.1 mD. Both dis-
tributions of ‘pore’ size are important. The attenuation Over-pressured zones due to rapid sedimentation of
can be shown to be the sum of Biot-type fluid flow and alternating sands and shaly sediments present a potential
squirt flow to/from the larger, moderately intercon- hazard when drilling, due to the risk of shallow water
nected inter-granular pores, which may contribute as flows (SWF). Effective stresses and compaction of sedi-
much as 90% of the total porosity. This permeability- ments can be minimal, and progressive instability during
Qp trend is expected to strengthen when the small-scale drilling at a new well can potentially engulf neighbour-
‘dual porosity’ also has the contribution of in situ jointing ing wells, also at depths up to a kilometre. Due to very low
or fracturing. A requisite number of joint sets for con- values of shear wave velocity at low effective stresses in
nectivity, and well-intersection for verification, are nec- sands, there is an exponential increase in the ratio of Vp/Vs
essary boundary conditions. to values beyond 5 and 10, and even beyond 100. Poisson’s
ratios increase rapidly to just below 0.5. There is high
attenuation of the shear waves at the lowest pressures, as
13.30 the sand is close to a state of suspension. Distinguishing
between unstable sand and sandstone is very clear using
Ultrasonic data (0.7–0.85 MHz) for small ‘intact’ dual- the ratio of Qp/Qs plotted versus (Vp/Vs)2.
porosity limestone samples has uncertain relevance for
geophysicists interpreting propagation through dual-
porosity porous and jointed limestones in the field, at the 13.33
lower frequencies used in seismic and sonic log surveys
(50 Hz to 30 kHz range). Dual-porosity chalk, with higher Saturated shales tested under over-pressured conditions,
porosity than limestone, has indicated Qp and Qs values show anisotropic velocity and attenuation in ultrasonic
594 Conclusions

laboratory testing. Velocities can be 30–40% higher par- in the direction perpendicular to the stress-induced
allel to bedding, where attenuation is also least. There is fracturing.
a general increase in seismic Qp and Qs with increasing
differential pressure. Relative proportions of Biot ‘fluid-
past-frame’ attenuation, and local squirt flow attenua- 13.36
tion are different in the plane parallel to the layering,
and in the plane perpendicular to the layering. There is The coupled stress-permeability-velocity behaviour of
a strong link between the rock framework, the pore smooth-planar and rough-undulating fractures is differ-
geometry and connectivity, and therefore of the ent. Rough fractures closing due to stress increase con-
response of pore fluid to the propagation of seismic tribute to increased velocity, but suffer less than expected
waves in specific directions. reduction in permeability. The reason for the different
behaviour of the rough fractures compared to the
smooth, may be that E (physical aperture)  e (hydraulic
13.34 aperture), for the case of rough, high JRC fractures (or
joints), while E  e for smooth fractures (or joints). This
The surface roughness of joints or fractures may help to would mean faster physical closure than hydraulic closure
maintain some permeability, even at higher confining for rough joints, thereby potentially explaining the
pressures, corresponding to depths of several kilometres. stronger velocity response and the weaker permeability
Permeability parallel to jointing or fracturing may then response to stress increase.
be much higher than that parallel to eventual sedimen-
tary layering. The additional possibility of pre-peak-
strength conjugate shearing of such joint sets, due to 13.37
anisotropic stress, would allow relative maintenance of
joint permeability despite high effective stresses. Elastic Reservoir-scale 4D seismic monitoring in fractured reser-
property anisotropy, and hydraulic anisotropy may be voirs is most sensitive to production-induced changes at
closely related in terms of symmetry directions, when the lower effective stress levels. The velocity and particularly
two mechanisms share the same cause, such as layering or the attenuation, are relatively sensitive indicators of small
jointing. permeability changes. By the nature of jointed reservoirs,
there are unlikely to be commercially viable hydrocarbon-
bearing fractures or joints with very low surface roughness
13.35 JRC values, as joint closure under stress would preclude
both permeability and ‘storage’, if such was needed due to
High pressure polyaxial loading frames have been used to low porosity matrix. Minerally ‘frozen’ stylolites in lime-
study the seismic signature differences between unfrac- stone and chalk are a special case, being both exceptionally
tured matrix and fractured matrix. Parallel fracturing has ‘rough’, and insensitive to stress change, in comparison to
been developed in the same specimen by holding the min- rough, interlocking joints which show greatest stress sensi-
imum principal stress very low and increasing 1 and tivity, and greatest apertures (E and e).
2 in unison, to high levels. Velocity increasing steadily
parallel to the high biaxial loading direction, contrasts
with the fall in velocity in the perpendicular direction, 13.38
when fracturing initiates. Subsequent reloading of the
fractured sample demonstrates stronger stress-velocity A simple empirical method is suggested for linking seis-
dependence perpendicular to the fractures than parallel. mic Q values, specifically Qp, with hydraulic and rock
Perpendicular to the fracturing direction, seismic Q val- engineering properties. This is based on the fact that Qseis
ues change from 40 to 30 as a result of fracturing, and invariably resembles the static E-modulus in the case of
reduce to 10 with unloading, actually resembling poten- intact samples, and the static deformation modulus in
tial deformation modulus behaviour, when expressed the case of jointed or fractured rock and rock masses.
as GPa. Low permeability sandstones start to develop when Qseis magnitudes are expressed in GPa. Despite the
measurable permeability when velocities start to reduce dynamic micro-strain basis for spectral analysis estimation
Conclusions 595

of Qseis, the magnitude of Qseis and its increase with often resulting in several MPa greater minimum stress
effective stress, does not so closely resemble the dynamic in the shale, which is frequently a fluid barrier for the
micro-strain based deformation properties such as Edyn hydrocarbon-bearing sandstone.
as one might expect. It has been found that Qseis cubed
and inverted gives first-order estimates of intrinsic per-
meability, and velocity and compression strength com- 14.3
bined (both reflecting stiffness), give independent
first-order estimates of Qseis. Fine layering of alternating porous and impermeable strata
is obviously one of the basic sedimentary systems that con-
tribute to the existence of potential reservoir rocks in sedi-
Chapter 14 P-waves for mentary basins. Fine layering of sedimentary strata means
characterizing fractured that the dominant wavelength of a seismic or sonic pulse
reservoirs is long compared to the thickness of individual layers. The
medium will nevertheless exhibit effective (and real)
14.1 anisotropy, with a vertical symmetry axis in the case of
horizontal layering. In the presence of hydrocarbons this
There was very early recognition in petroleum explo- layered medium may show substantial attenuation and
ration, of velocity increase with depth, and early recogni- velocity dispersion, which will be compounded with the
tion of a quite systematic trend linking velocity to the additional presence of jointing or fracturing. With mod-
geological age, in combination with the present depth of ern seismic techniques a new exploration concept has
occurrence. The greatest rate of velocity increase was gradually developed, exploring not just for the presence of
found to occur at shallow depth in the oldest units, reservoir rock containing hydrocarbons, but exploring for
which is fundamental early proof of the importance of the presence of permeable joint-sets and their principal
dual porosity. The likelihood of more joints in the stiffer, direction.
older units means that these units are more sensitive to
stress change. However, with only Vp as a dynamic indi-
cator of conditions, acoustic closure represents a limit to 14.4
the stress-sensitivity of velocity, especially for the case of
weaker, younger reservoir rocks. A ‘thin bed’ is considered to be 3/8 of a wave length, the
limit for a discrete reflection both from the top and bot-
tom of the bed. Wave scattering, attenuation and disper-
14.2 sion occur when the ordered heterogeneities have scale
lengths of about 0.3–0.01 of the wavelength, while the
An early analysis of almost 300 kilometres of well sec- smallest scale of ordered heterogeneity, less than 0.01
tions, in 500 petroleum well surveys, mostly from the of the wavelengths, may be the cause of most of the
USA, and mostly for mixed shale and sandstone sec- azimuthal and offset dependent velocity. Conventional
tions, indicated an average P-wave velocity discrepancy seismic wavelengths are much longer than the scale
of only about 110 m/s in velocity between these two, lengths of either of the features that govern dual-porosity
the sandstone having the highest velocity by this small flow in a reservoir.
average margin. The similarity of velocities for these
mechanically and hydraulically dissimilar units, is a
reminder of the potential ‘non-uniqueness’ of P-wave 14.5
velocity, and the need for alternative interpretation meth-
ods, such as attenuation, and impedence, to distinguish Strong P-wave velocity anisotropy is observed in every
the different lithologies and their fluid-bearing signa- geologic environment, with the possible exception of
tures. The closeness of the in situ velocities for shale and basins under primary deposition and burial. P-wave
sandstone, also seen in rock physics experiments on the azimuthal anisotropy, previously ignored and left to the
matrix of both rocks, is surprising, in view of the greater research and technology specialists, is now known to be
tolerance of the stronger sandstone to stress anisotropy, one of the most significant properties of the acquired
596 Conclusions

seismic data. In the marine environment, fully populated Qseis/Vp ratios could therefore be used to delineate the dif-
offsets in each azimuth bin are less common than on land, ference between unconsolidated sands and jointed sand-
but even narrow azimuth data gives an opportunity to see stones, and between weak plastic shales, and the less
the effects of azimuthal anisotropy. As time goes by more desirable fissured/jointed, or indurated variety.
and more reservoirs are being re-classified as severely het-
erogeneous, as well as fractured and anisotropic.
14.9

14.6 At a site in the USA, the P-wave velocity of near-surface


jointed limestones determined from near-offset VSP var-
Definitions of three common classes of anisotropy are as ied from about 3.8 to 4.2 km/s between the shallow
follows. Transversely isotropic media have a vertical axis depths of 16 and 26 m. This suggests a rock quality
of symmetry, and are referred to as TIV: with horizontal Q-value of 2 to 5 from the empirical relation Vp 
axis of symmetry as TIH. The former is typified by fine 3.5  log Q relevant to nominal 25 m depth, 100 MPa,
layering in shales. Transversely isotropic media with a low-porosity rock. Qrock  2 to 5 is typical for rock
horizontal axis of symmetry, known as HIV, typically masses with three sets of joints, moderate block size, and
have stress-aligned, vertical jointing or fracturing, and/or with possible weathering of the joint walls: i.e. Q 
microcracks. When fast and slow directions have been 90/9  2/2  0.66/2.5  2 to 3. A seismic Qp value of
identified, azimuth sectoring can be applied in these about 13–14 can be estimated via the deformation mod-
directions. Even isotropic processing codes can function ulus method, when Qrock  2 to 3, and assuming the
with such azimuth-sectored data. uniaxial compressive strength of the limestone is around
100 MPa. Alternatively, using just velocity, Qseis can be
estimated as 12–16 for this jointed limestone, based on
14.7 the Vp range of 3.8–4.2 km/s.

Shallow 3D-survey based imaging of seismic velocity


and attenuation, shot over several km2, can be used to 14.10
derive tomographic images of velocity and attenuation at
specific depths in shallow reservoirs, using an iterative In this ‘circular-logic’ jointed rock prediction, the ratio
reconstruction algorithm. The tomograms derived for Qp/Vp, if roughly correct, would be a much lower 3 to 4,
constant-thickness slices centred at increasing depths compared to the much higher ratios seen in porous,
show successive increases in both P-wave velocity and unconsolidated sediments. A 460 m deep cross-well meas-
Qp with increasing depth (e.g. Qseis  5 when Vp  urement in a very permeable limestone aquifer showed Vp
2 km/s, and Qseis  10 when Vp  3.5 km/s). and Qp of 3.5 km/s and 14 at 2 kHz, but at higher fre-
quencies (12 kHz), Qp rose by 350%, while Vp rose by
only 3%.
14.8

Smaller scale, shallow cross-well tomography, has been 14.11


used in reservoir sands (channel-sands) to correlate high
values of the ratio Qseis /Vp with the most porous and most There have been many years of oil-industry interest in
permeable zones, and low values with flow barriers, such cross-hole tomographic methods for imaging below the
as shale-rich layers (e.g. sands Qp/Vp  45/3  15, plas- resolution of surface seismic. High frequency waves
tic shale Qp/Vp  30/3.7  8). Absence of jointing and can be propagated over distances of many hundreds of
high porosity and permeability apparently causes higher metres, with minimum loss of energy, when both source
values of Qp (i.e. less attenuation without squirt losses), and receiver are down-hole, in deep boreholes. The
and quite low Vp. This is in contrast to the seismic attrib- avoidance of near-surface (low Qseis) losses, means that
utes of jointed sandstones or jointed rock in general. The broad band-widths can be used. Very high resolution
proposed Vp-UCS-deformation-modulus ‘GPa-model’ images are obtained by using second arrivals and reflec-
for Qp fits the latter and not the former. A contrast in tion imaging.
Conclusions 597

14.12 include azimuthal variation of attenuation in 4D moni-


toring, possibly due to conjugate joint shearing.
At a 260 m deep research well through finely inter-lay-
ered limestone, shale, and sandstone sequences, the com-
bined use of a borehole compensated sonic logging tool, 14.15
a compensated formation density tool, and a formation
micro-scanner enabled resolution of much of the detail Since all seismic data also varies with frequency, there is
of finely interlayered rock sequences, where the normally increasing acceptance that 3D multi-component, multi-
detectable layer thickness using standard sonic tools may mode and multi-azimuth acquisition may be required
be no less than 15 cm. The ability to separate shales, and may also be economically justified. A given set of ver-
sandstones and limestones, based on down-hole facies tically aligned fractures may cause anisotropy with low
recognition and velocity differences, can also be used in a frequency measurement, signal distortion with mid-fre-
tentative separation of Qp according to facies. With quency measurement, and lead to reflections by the high-
appropriate ranges of increasing uniaxial strengths and est frequency waves. The detection of the azimuthal
velocities for the three rock types shale, sandstone, lime- anisotropy attributable to structure, that specifically con-
stone, one arrives at potentially representative Qp values trols the fluid-flow properties at reservoir scale, is now an
of 6–7, 10–15, and 20–40, using the empirical Vp-UCS- important focus of attention. The detection of permeabil-
modulus model. ity anisotropy can be considered as one step beyond the
detection of vertically aligned fractures, and/or the detec-
tion of unequal horizontal stresses.

14.13
14.16
A finely inter-bedded mix of facies as above, may tend to
create a ‘weighted’ response in standard logging. Sonic log- There is also a growing trend to instrument selected petro-
ging (8–24 kHz) gave the lowest Qp with a mean Qp of 10 leum wells on a permanent basis, especially offshore, so
and a range of about 6–14. A Qp range estimated from the that 4D seismic can be used relatively more easily, to mon-
measured velocity range of 3–4 km/s, using the Vp-UCS- itor changes bought about by different water-flood and
modulus method, would be about 5 to 12. The somewhat production practices. Rock physics principles are used to
lower frequencies of cross-hole logging (0.2–2.3 kHz), assist in the interpretation of measured changes in veloc-
gave a mean Qp of 15.7 and a range of 12 to 20. The low- ity, amplitude and attenuation. In-well 3D accelerometer
est frequency VSP (30–280 Hz), giving presumably the installations were applied in the late 1980’s for permanent
poorest definition of the fine inter-layering, gave a mean installations in deep holes adjacent to the San Andreas
Qp of 31.3 and a range of 25 to 45. fault in California, where the benefit of avoiding near-sur-
face attenuation were recognised.

14.14 14.17

Since sedimentary rocks containing hydrocarbons have It is commonly assumed that there is strong correlation
proved to be neither isotropic nor homogeneous, but het- between directionality of reservoir flow and the local,
erogeneous and anisotropic, the seismic wavelength at presentday orientation of the maximum horizontal
which the measurement is made, determines what seis- stress. Oriented four-arm calliper logs typically show a
mic attributes can be measured, and whether the rock long axis that is oriented parallel to the minimum hori-
looks homogeneous and isotropic or heterogeneous and zontal stress direction, if there is stress-induced break-out.
anisotropic. All seismic data are now known to vary with However there may be geomechanics-based reasons for
offset from the well (in VSP) and with azimuth. Besides carefully evaluating this commonly held viewpoint from
detecting azimuthal velocity anisotropy due to aligned case to case. Rock strength, joint or fracture roughness,
fracturing or stress, one can now acquire spatial resolu- joint closure under stress, and possible shear-displacement
tion of variable structure, azimuthal resolution of attenu- modes need also to be considered. Fractures perpendicular
ation, and resolution of temporal changes, which may to the H max direction can also be ‘open’ if partially filled
598 Conclusions

with mineral cements, and for this same reason, sealed fluctuations. In addition, minimum rock stress will tend
fractures parallel to H max are also numerous. to be residing in the weaker beds (i.e. shale) at shallow
depth, while residing in the stiffer beds (i.e. sandstone) at
depths where shale is more plastic, and therefore has a
14.18
higher h min. Stress concentrations around wells will
cause a magnification of tangential stress, in the same
The common assumption that the direction of H max is
direction as major principal stress, and diminution of tan-
the direction of ‘open’ cracks or fractures also overlooks
gential stress in the perpendicular direction. If these stress
the possibility that two sets of joints or fractures can be
concentration effects are strong enough in relation to rock
involved in limited conjugate shear-displacement. With
strength, shear failure surfaces may develop, first giving
one set dominant, the orientation ‘discrepancy’ often
break-out, subsequently a possible log-spiral-sheared dis-
reported, regarding the direction of ‘open’ fractures in
continuum close to the well.
relation to H max, can be better explained. Dominant
directions of fracturing can also be images of the result-
ing dominant strike of conjugate, steeply dipping sets, in
14.22
the case of domal structures.
Seismic attenuation has come to be recognised as poten-
14.19 tially very sensitive to reservoir properties. This is because
of its sensitivity to fractures, joints or bedding planes, and
A further potential source for minor angular discrepancies, in turn, due to their sensitivity to changes of effective
is the dilation-related contrary-rotation of fluid lenses con- stress and to frequency. Attenuation levels are also sensi-
tra rock-to-rock contacting asperities, when non-planar tive to the saturating fluid and petro-physical properties.
joints or fractures are under significant shear stress, and High dispersion (and low Qseis) values may correlate with
therefore significantly ‘open’. This geometric effect could permeable sand and carbonate beds within shale. Such
also potentially cause a minor rotation of shear wave split- beds can be at least ten times as permeable as the host
ting polarization, as argued in Chapter 15. shale formation. The dependence of seismic velocity on
frequency can be used for reservoir characterisation, since
the dispersion is mathematically related to seismic attenu-
14.20 ation. High frequency measurements differ from low fre-
quency measurement due to both elastic scattering and
Measurements in deeper wells have indicated that seismic intrinsic attenuation.
Qp based on seismic frequency VSP, may be systematically
smaller than Qp based on higher frequency sonic logging.
This is the opposite of what has been measured in shal-
14.23
lower wells, where attenuation was least for VSP and most
for sonic logging, giving lower Qseis. The expected disper-
AVO (amplitude variation with offset) and AVOA (ampli-
sive bias of higher frequency (sonic) waves travelling at
tude variation with offset and azimuth) indicate that
higher velocities than lower frequency seismic (VSP)
variation of P-wave amplitude can be related both theo-
waves, remains consistent when shallow or deep.
retically and in practice, to the presence of fracturing.
Individual values of Qp or Qs may change ‘erratically’ with
Appropriate analysis of AVOA gives reasonable estimates
depth unless depth averaging is used. However, rock qual-
of the orientation of fracturing, particularly if only one set
ity Q-values down recovered core also tend to fluctuate
is involved, or if one set is dominant. Fracture orientations
quite strongly, and since linked to deformation modulus,
can be compared to results obtained when using C-waves
Qp and Qs must also be expected to fluctuate.
(P to S converted waves), and the shear-wave splitting and
polarization mechanism. The converted P to S waves have
14.21 the advantage that they can be generated by compres-
sional (i.e. explosive) sources, yet are expected to contain
Rock quality differences, and therefore differences in stiff- the same information as pure S (or SS) waves, as discussed
ness and susceptibility to failure may play a role in such in Chapter 15.
Conclusions 599

14.24 14.27

Although the use of shear waves are theoretically There is a reported problem of model-dependence in
favoured for fracture set detection, there has been some AVAO analyses. The dominant fracture strike direction
reluctance to use shear waves, due to more expensive can be ambiguous, since the azimuthal variation in the
acquisition and more expensive processing routines. near-offset AVO gradient, can be positive or negative, rela-
For these reasons, the use of P-waves for fracture set tive to the fracture direction. The direction of the most
detection and estimation of orientation, has attracted a positive AVO gradient can correspond to either the frac-
lot of interest, even though P-wave travel times need to ture-normal, or the fracture strike direction, depending
be detected in many directions to obtain the necessary on the character of the fracturing, and depending on
information. whether brine-filled or gas-filled. Forward modelling is
therefore needed in order to constrain the interpretation
of AVOA. Forward modelling using fracture density, frac-
ture aspect ratio, and fracture (additional) compliance con-
14.25 cepts, may actually require knowledge of two different
fracture apertures: the hydraulic aperture (e) that would
If seismic data acquisition is conducted parallel to the govern squirt losses, and the physical aperture (E) that
(geologically suspected) fracture orientation, the fractures would govern compliance or stiffness and stored volume
will have minimal influence on the reflection properties, of fluid, where E  e. This inequality in the case of
regardless of the angle of incidence, or offset. The P-wave favourably rough-walled joints or fractures seems so far to
particle motion is then parallel to the fractures. If the seis- have been ignored.
mic line is instead oriented more perpendicular to the
fractures, at larger angles of incidence than zero, the reflec-
tion coefficients will be affected strongly. At the largest
14.28
angles of incidence, especially perpendicular, the P-wave
velocity is also expected to be affected by the acoustic
The probability for multiple joint or fracture directions in
properties of the fluids filling the fractures. Thus in the
the neighbourhood of faults, means that the fast velocity
presence of anisotropy, the reflection amplitude will vary
is no longer equal to the matrix or bulk rock velocity.
with offset, due to changed angle of incidence, and will
The normal elliptical Vfast and Vslow distribution is then
also change with azimuth (AVOA).
replaced by superimposed multiple ellipses, which have
the effect of reducing the observed velocity. The previous
directionality with a single set of joints or fractures will be
14.26 lost. Due to rapid changes in fracture frequency, rapid
changes in velocity are also seen. Such is actually a
When deviation of 20° to even 40° is observed between response to the rapid changes in rock mass quality Q close
AVO-determined dominant fracture orientation and to, and across faults, as frequently mapped in tunnelling,
the perpendicular-to-break-out based H max direction, and when logging fault-zone core, in each case for rock
and when nearly as large deviation is also obtained quality determination.
between shear-wave polarization and the H max direction,
the possibility of conjugate-shearing of joint sets that are
intersected by the H max direction should be considered. 14.29
Although this contradicts the standard industry assump-
tion of ‘open fractures parallel to H max’, it helps to There is a multitude of technical jargon in the geophysical
explain frequent angular discrepancies between domi- industry. Some is exceedingly simple. Converted C waves
nant ‘open’ fracture azimuths, and the perpendicular-to- means explosive or air gun generated P-waves converted to
break-out based H max direction, At shallow depth, the S-waves at an interface or at the sea floor. (Pure S-waves
standard industry assumption, also in civil engineering, may be referred to as SS). The term 4C means four-
is more correctly focussed on maximum permeability, component seismic recordings. These consist of one
and Vp, being parallel to the H max direction. hydrophone, one vertical geophone, one in-line horizontal
600 Conclusions

geophone, and one cross-line horizontal geophone. The compaction, and reduced layer thickness, i.e. reinforc-
term 4D means 3D seismic repeated at intervals for mon- ing effects. Time-shifts of as much as 12–16 ms, between
itoring changes caused by production. Repeated 4D sur- 1989 and 1999, recorded at the Ekofisk field, were related
veys can be made cheaper by modifying 4D receivers to to an estimated 6m of additional compaction at 3 km
also be 4C. There are now some expensive 4D4C installa- depth. An uncritical time-lapse comparison between sur-
tions that make frequent reservoir monitoring much veys, may give unrealistically large values for compaction
cheaper. The consequence of frequent full-field 4D4C re- and subsidence.
shoots, as at Valhall, in the North Sea, providing full-field
estimates of all required reservoir parameters, is more effi-
cient exploitation of reserves, and a production increment 14.33
obviously coming sooner than discovery and exploitation
of new fields. Extensive casing damage to numerous wells at Ekofisk is
evidence of discontinuous behaviour, due to stretching of
the overburden and differential bedding plane slip. This
14.30 was also seen in early discontinuum models. Subtle
changes are now known to occur to the overburden
Reservoir monitoring with 4D seismic in its most basic velocity, due to the ‘stretching’ of the overburden in
form is the repeated inversion of changing seismic data, to response to the incremental compaction between surveys.
obtain dynamic reservoir properties, which can subse- There are about 150 km3 of obviously discontinuous rock
quently be used to predict pore pressure change at a dis- involved in the compaction and subsidence. Further evi-
tance from the wells based on the effective stress and fluid dence for discontinuous behaviour caused by compaction
sensitivity of laboratory samples of the reservoir rocks. at Ekofisk can be seen in the results of the 4D seismic.
History matching can be used to up-scale the rock physics Fault related discontinuities are seen in ‘time lapse’ tomo-
matrix data, and to calibrate forward modelling of grams of compaction magnitudes. Forward modelling of
anisotropy and fracture effects. There is high sensitivity to compaction details, performed in the 1980s, indicated
effective stress in shallow reservoirs, but a stress-velocity small-scale down-dip shearing of conjugate jointing in the
plateau may be reached at high effective stresses (i.e. chalk, before evidence of slickensiding had been seen in
beyond roughly 25 to 50 MPa, depending on rock type newly drilled core.
and on fracture-surface roughness.) A velocity plateau
indicates the need for up-scaling using attenuation.
14.34

14.31 In water-flooding, for stimulating and driving petro-


leum production, there is both a local increase in pore
Compressible grain-boundary cracks with their low pressure at the injector wells, and a reduction in tempera-
aspect ratios may be partly the result of stress unloading ture, causing some contraction of the matrix, both of
when matrix samples are drilled and bought to the which help to dilate, and possibly shear existing joints,
surface. The in situ velocity may not be recovered upon and perhaps create new fractures as well. The conven-
reloading, due to hysteresis caused also by temperature tional and expected mechanism of fracture or fault
change. Saturated samples containing micro-cracks will opening exactly in the direction of Sh min is not as com-
tend to project a lower stress sensitivity with laboratory mon as expected.
ultrasonics, than with seismic waves, due to greater
relaxation, as opposed to stiffening with the ultrasonics.
14.35

14.32 There are strong indications from surveys of numerous


fractured reservoirs that have been water-flooded, that the
When compacting reservoirs are 4D monitored by reduction of effective stress caused by the water-flood pres-
repeated seismic surveys, time-shifts are registered, which sure, and the related contraction-cooling effects, could
are a combined result of increased velocity due to be stimulating shear-displacements on existing joint or
Conclusions 601

fracture sets, or faults. The direction of H max shows a fre- are rough, or that there exists a close-to-fracturing pore
quent tendency to have bisected the geologic features that pressure, or that there is a suitable quantity of hard min-
are the basis for the joint rosettes. An implied mechanism eralization to ‘bridge’ and maintain an earlier porosity.
of conjugate shear, would show strong parallels to the Alternatively, if some pre-peak-strength shear displace-
findings of Zoback and co-workers, concerning the fre- ment of non-planar joints or fractures has occurred,
quency of water conducting discontinuities in deep wells there would be the contribution of dilation to ‘open-
needing to be oriented so that they are under shear stress. ness’, and the additional influence of a 10° to 20° rota-
tion of the fluid-bearing parts of the fractures in
relation to the contacting parts taking the load. This
14.36 might rotate both the sources of attenuation and the
sources of shear-wave polarization.
Mapping of the azimuthal velocity anisotropy of P-waves
using a downhole triaxial accelerometer sensor array, and
multi-azimuth walk-away (or ‘float-away’) VSP, is a means Chapter 15 Shear wave splitting in
of reducing the risk of drilling low-productivity wells in fractured reservoirs and
unfractured parts of reservoirs. Calibration with oriented resulting from
core data and with FMS logs improves the likelihood that earthquakes
later producing-wells will intersect ‘open’ conductive frac-
ture sets, where P-wave anisotropy is highest. Azimuthal 15.1
variation in the shear modulus of the fractured rocks is
cited as the reason for the P-wave velocity anisotropy. Vertical and sub-vertical jointing is extremely common
Dominance of one fracture set orientation, with a near- in most rock masses. Yet vertical boreholes are usually
orthogonal subset, and variation of fracture density in the the first, and seemingly also the second choice, for sam-
unequal two-set system, would be reasons for variations in pling and gaining access to the sub-surface. The sam-
the degree of seismic anisotropy, and demonstrate the pling bias caused by the mismatch of borehole diameter
benefit of 3D mapping. and horizontal spacing of vertical structure, and the
vertical borehole itself, is extreme and well known. If
the economic savings of a vertical well, and the subse-
quent cost of an extensive seismic survey and its inver-
14.37 sion were combined, there would perhaps be reason for
rapidly deviating exploration boreholes at least 10° or
Multi-azimuth walk-away VSP can also be used to map 15°, in order to sample the increasingly understood rele-
the attenuation anisotropy of a reservoir. A fractured, oil- vance of vertical and sub-vertical structure on hydrocar-
saturated reservoir is likely to show azimuthal variation in bon production. On the other hand drilling and hole
attenuation, in a similar manner to P-wave anisotropy. As stability problems might be increased by the more fre-
examples, we may quote Qp  18 in the fractured part of quent joint intersections.
a reservoir, and Qp  35 to 40 in the overburden, which
was assumed to be relatively unfractured, with minimum
attenuation correspondingly scattered between wider 15.2
azimuths. In the particular reservoir, minimum attenua-
tion was some 20° to 30° oblique to both the open con- Shear-wave anisotropy due to splitting and polarization
ducting fractures and to the h max direction, based on caused by the presence of vertical or aligned structure,
oriented cores and borehole images. Conjugate jointing, (and P-wave azimuthal anisotropy), are miraculous means,
perhaps also pre-peak shearing, is again suggested. in view of the long seismic wave lengths, for rectifying
these poor joint or fracture sampling strategies. The cen-
tral challenge of sub-surface fracture characterization is to
14.38 obtain data on essential fracture attributes where direct
observation has been prejudiced by vertical wells. An early
‘Open’ fractures in a petroleum reservoir would seem to deviation of 10° or more would greatly improve under-
require that the rock is unusually strong and that joints standing of both the overburden jointing and the reservoir
602 Conclusions

jointing. The potential anisotropy of the overburden can- fluids prop open a population of compliant voids or
not be ignored in seismic inversion, as indicated in an inclusions that are nevertheless capable of remaining open
increasing number of cases, especially where compaction against the least principal stress. The implication of 3D
and therefore subsidence are occurring. principal stress anisotropy at depth is that EDA cracks
will tend to be aligned in a vertical plane, striking parallel
to the major horizontal stress. With this configuration, a
15.3 microcracked but otherwise isotropic crust would be
transversely isotropic, with a horizontal symmetry axis.
Countless hydrocarbon reservoirs have been discovered,
characterised, and monitored by P-waves. However,
P-waves cannot solve every seismic imaging or reservoir 15.6
description problem. The addition of S-waves, usually in
the form of converted PS-waves, has given oil and gas The traditional view was that there were several possible
companies an enormous quantity of new reserves that small scales of azimuthal anisotropy that could cause shear
could not have been found with P-waves alone. The new wave splitting, such as aligned crystals, lithological
reserves have been more effectively exploited by better anisotropy due to aligned grains, stress-aligned microc-
identification of fracturing, and therefore better place- racks, and fine layering. Much evidence for the influence
ment and deviation of production and water-flood wells. of larger-scale joint-set alignment effects on shear wave
Multi-component recording of shear-wave attributes also splitting has subsequently been obtained. These larger
provides information where shallow gas has obscured P- scale features obviously dominate drainage potential from
wave imaging over central parts of a field, such as at the matrix to the joints, and thence to the wells in hydro-
Ekofisk and Vallhall. carbon production. The micro-scale extensive dilatancy
anisotropy (EDA) championed by Crampin and co-work-
ers, would logically dominate drainage from the pores to
15.4 the microcracks. Relatively large time delays between split
shear-waves may also be set up in the top tens to hundreds
The basic geological source of polarized shear waves can of meters of rock. These near-surface effects have been
be sets of vertical joints or fractures, or stress-aligned termed natural directivity. Principal stress aligned microc-
microcracks. These aligned features cause the vertically racks, or principal stress aligned intra-bed jointing, or
transmitted shear-waves to split into fast and slow compo- aligned jointing from historic tectonic effects including
nents, registered as time delay, due to the attenuating doming and anticlines, each have shear-wave splitting
effect of fracture shear compliance, on the S-wave compo- potential.
nent that has particle motion perpendicular to the fracture
strike. Shear-waves travelling in the parallel direction
hardly sense the presence of the cracks, and travel at 15.7
almost the wave speed of the unfractured matrix. The dif-
ference in travel-time between the fast qS1 and slow qS2 It was earlier considered remarkable that, with all the dif-
components is strongly related to the length of travel path ferent scales and characters of aligned fluid-filled cracks,
and to the density of the crack population. It is also related inclusions or fractures in sedimentary, metamorphic and
to fracture compliance. Numerous polarized shear wave igneous rocks, that the differential shear-wave anisotropy
observations show the fast wave polarized parallel (or sub- varied only within narrow limits (0.5 to 5%). With
parallel), to the accepted local or regional maximum stress increasing application at fractured reservoirs, this range,
field. There can be several reasons for this, and also several and the earlier assumed limited ranges of fracture den-
reasons for deviation from this direction in other cases. sity, have each been extended, sometimes by significant
margins.

15.5
15.8
The potential of fluid-filled microcracks to react to crustal
stress and strain led to the early proposal for extensive dila- The traditional porous medium experience is that shear-
tancy anisotropy (EDA). The hypothesis was that crustal wave velocity remains unchanged whether a formation
Conclusions 603

contains gas, oil or water. However, because of the effects are sure to vary, and each split shear wave may therefore
of fluid compressibility on the dynamic normal stiffness split again, giving multiple splitting. The influence of
of fractures, or its inverse compliance, the polarized shear the joint structures near the recording site gives one of
waves passing through a fractured or jointed medium, the most prominent results. Down-hole instrumenta-
may actually have the unexpected ability to distinguish tion is needed if possible, in order also to minimise
between oil and gas, specifically when incident waves and such site effects and the higher frequency filtering due to
jointing are non-parallel. Regions of gas are characterized attenuation.
by lower magnitude slow shear waves, and regions of oil by
higher magnitude slow shear waves. In comparison, clas-
sic Gassmann porous medium theory anticipates an
15.12
S-wave velocity relatively unaffected by the type of fluid.
High velocities in rock tend to occur where attenuation is
low or Qseis is high. This reciprocal relationship between
15.9
velocity and attenuation is one of the reasons why the
leading split shear wave is a very stable phenomenon,
A controversial point is whether the fractures and shear-
because it is travelling in the fast direction and is less atten-
waves can both be vertical, where theoretically only shear
uated than the slower split shear wave. Sometimes the
compliance would be sensed, because it is uncertain if
slow wave is too attenuated to allow calculation of the
shear compliance will be sufficiently affected by these
shear-wave anisotropy. The location along the ray path
contrasting fluid compressibility effects. A more certain
where the shear-wave splitting is imprinted most strongly
effect of fluid compressibility is when normal compli-
is not known a priori, and near-surface effects where joint-
ance is involved in the case of sub-vertical fractures or
ing is stronger may be a disturbing feature. Well fractured
sub-vertical shear-waves, giving a finite angle of inci-
reservoirs may over-print such effects. Lower crack densi-
dence. There is evidence that the delay between the split
ties implied by many earthquake studies may be a reflec-
shear waves may decrease with increasing depth, yet an
tion of the sampling of ‘average rock’. In contrast, crack
accumulative delay with increasing depth would naturally
densities interpreted from fractured reservoirs may repre-
be expected. This perhaps suggests that stress-sensitive
sent a ‘biased sample’, caused by a rock mass that is more
compliances are involved, which would match rock
jointed or fractured than the norm, and therefore also a
mechanics experience with the non-linear pseudo-static
target for exploration and subsequent exploitation.
stiffness of joints or fractures.

15.10 15.13

In the case of earthquake recordings, the relative steep- The geophysicist’s crack density (e) is defined as number
ness required for the incident wave to make an acute (N) of cracks per volume (V) times the crack radius (a)
angle to typical sub-vertical structure, means that there is cubed. Crack density was often quoted in the range of
a need for the recording site to be within the shear wave e  0.01 to 0.05 in reportedly widely different geologi-
window. This derives from the requirement of angles of cal and tectonic regions. This commonly used parame-
incidence less than sin1(Vp/Vs). For a Poisson’s ter is unfortunately remarkably ambiguous. Ten million
ratio of 0.25, is about 35°. Outside this window the micro-cracks @ 100
m/10 cm cube, give e  0.01,
shear-wave waveforms are severely distorted. The epicen- while ten fractures @ 1 m/10 m cube also give e  0.01,
tral distance from the recording sites must therefore be and even ten minor faults @ 100 m/1 km3 give
significantly less than the focal depth of the earthquake. e  0.01. These three scenarios, with their theoretically
equal ‘crack density’, have very different mechanical and
fluid-conducting properties. Nevertheless they theoreti-
15.11 cally would previously have suggested equal shear wave
anisotropy. Alternative methods of forward modelling
The large depth of most earthquake sources means that show this theory to be in error, when extremes of frac-
shear waves will pass through a range of rock types with ture size and compliance are involved, and due to fre-
different ages. Velocity and individual joint-set properties quency or dispersive effects.
604 Conclusions

15.14 Shear wave splitting was also visible in reflections from


layers above the reservoirs. More recently, shear wave
The classically assumed limited range of 1% to 5% shear- splitting has been observed to mimic saucer-shaped and
wave anisotropy, was linked to a similarly limited range of oval-shaped sub sea-bed subsidence bowls kilometres
assumed crack densities 0.01  e  0.05. The percent- above compacting jointed reservoirs. Strength-scaled
age of differential shear wave anisotropy was reportedly polarization patterns show excellent correlation to loca-
about e  100, for a Vp/Vs ratio of about 1.7 or 3. tions where sea depths are changing most rapidly. The
There are now known to be many exceptions to these amount of detected anisotropy is small above the centre
limited levels of shear-wave anisotropy, and there are of the compacting field, where the subsidence is largest,
order of magnitude larger values of fracture density inter- and large on the flanks, becoming small again further
preted in fractured reservoirs. So-called fracture critical- from the centre. Oriented joint-stretch in the overbur-
ity, associated with higher densities, is clearly positive for den seems likely to be the cause, as a partially cubic
good reservoir production. The rock mass does not frag- polarization pattern resembling intra-bed jointing can
ment, and the shear strength is not lost, and nor does the be noted. If unconsolidated un-jointed sediments were
pore fluid disperse at higher crack densities, as feared by a the actual source of this nearer-the-surface splitting and
prominent author. This is because it is confined by a 3D polarization, then micro-cracks or even macro-cracks in
stress field, and by less permeable boundaries. How-ever the sediments would need to be invoked to explain the
rock mass deformability and susceptibility to com- polarization match to the subsidence bowls.
paction may obviously increase when looking beyond
matrix compaction mechanisms, as at Ekofisk.
15.17

15.15 On occasion, ‘90°-flips’ in polarization directions are


observed from presumed earthquake source zones. It
Aligned fracturing may be detected and monitored over a has been postulated that this may be due to extreme
huge range of length scales, using polarized shear waves. build-up of pore pressure, causing the faster split shear-
Dimensions may range from crustal dimensions of waves that were previously parallel to H max to do a
10–100 km, through 1–1000 m reservoir scale fractures 90°-flip and become the slower wave parallel to h min.
and faulting, to millimetre and micron-sized microc- From a geotechnical viewpoint a 90°-flip in polarization
racks. The relative stiffness of microcracks, having much would appear more likely with extreme H max loading,
higher aspect ratios than inter-locked fractures or joints, causing lateral expansion of aligned microcracks in a lim-
means that they cannot respond in the same way as frac- ited volume of rock. The volume affected by expanding,
tures, to a given change in fluid pressure, according to aligned microcracks may need to be limited, since the
classic geophysics teaching. Recent poro-elastic fluid rock mass could not absorb this volume increase without
interaction modelling of the dispersive effects caused by a general reversal of the H max and h min directions.
fractures of widely different dimensions, using double-
porosity or triple-porosity models, show that different
fracture dimensions can be inverted from given levels of
15.18
shear-wave anisotropy, based on their different response
to changing frequency.
The progression from seismic propagation in isotropic
media, to anisotropic layered media, to transversely
isotropic layered media containing one set of vertical
15.16 fractures, later increased to two sets of perpendicular
fractures, then non-orthogonal vertical sets, and finally
By the mid-eighties, some oil companies were appar- to non-vertical sets, has resulted in a progression of theo-
ently reporting shear wave splitting in almost all their retical papers in the geophysics literature, containing an
three-component reflection surveys in sedimentary increasing content and complexity of 6  6 compliance
basins. The phenomenon was assumed to be due to and stiffness matrices. Schoenberg and Sayers are promi-
fracture or joint sets within the fractured reservoirs. nent authors.
Conclusions 605

15.19 media, in which the shear modulus should be independ-


ent of the fluid. In the case of fractured media, oil and gas
The elastic moduli and the density determine the behav- are distinguishable by the reduced and increased shear
iour of seismic waves, assuming a linear, loss-free, elastic wave anisotropy respectively, as the stiffening effect of
behaviour. The presence of fracture sets affects the elastic the oil makes the fracture normal stiffness less contrasted
moduli of the fractured rock, due to the addition of their to the back-ground medium. So for dipping joints or
dynamic compliance. The additional presence of frac- fractures, there proves to be a significant decrease in shear
tures can be expressed as the sum of the compliance of wave anisotropy if the fluid has a higher bulk modulus,
the isotropic back-ground rock and the excess compliance making the normal stiffness of the fractures greater. The
matrix associated with the fractures. The latter is com- average of the two shear wave velocities is therefore also
posed of the effects of a fracture-normal compliance ZN, increased.
and of a fracture-shear compliance ZT.

15.22
15.20
The shear compliance ZT and normal compliance ZN
The simple addition of the three fracture compliance interpreted from loaded, roughened Lucite (Plexiglas)
terms (ZN , ZT and ZT) is made in the same diagonal-term plates, which were used to simulate ‘a fractured medium’,
(1,1 5,5 and 6,6) locations in the combined compliance was apparently responsible for some authors to assume
matrix. In the context of shear-wave splitting, the stiffness that ZN  ZT for the case of dry, gas saturated cracks.
matrix term relating to the fast shear wave propagating Seismic phenomena observed in highly stressed, finely
parallel to the fractures, is the C44 term, and the slow shear layered (t 0.7 mm), roughened plates of Lucite, with
wave propagating perpendicular to the fractures, is given their extreme ‘crack densities’ and artificial ‘fracture’ sur-
by the C55 term. The Thomsen shear-wave splitting faces, should however not be used to predict rock joint
anisotropy parameter () is defined from the elastic stiffness response to dynamic or static loading. The suggested
matrix as (C44 – C55)/2C55 (often expressed as a percent- ‘equality’ of ZN and ZT was propagated in some of the
age), where qS1  (C44/), and qS2  (C55/). Shear- geophysics literature, but may be far from realistic for all
wave splitting anisotropy as defined by Thomsen, is but the smallest laboratory specimens. ZN involves
commonly in the range 0    20%. The Crampin ‘micro-closure’ in a stiffening direction, while ZT
definition of the S-wave velocity anisotropy is a contrast- involves ‘micro-slip’ in a direction that may not involve
ing 100  (Vs max – Vs min)/Vs max or 100  (qS1 – qS2)/ stiffening.
qS1. For the case of the vertically propagating waves
through the vertical fractures, there is no fracture compli-
ance term in C44, only the Lamé constant
for the back- 15.23
ground rock. This causes the fast shear-wave component
to be parallel to the fractures. For the case of the slow shear An inequality of the joint or fracture compliances would
wave, the more complex C55 term involves ZT, and be more consistent with the experience of Ks  Kn, con-
not ZN. The simpler compliance matrix S55 term is simply cerning the pseudo-static shear and normal stiffnesses of
1/
 ZT. joints and fractures, where stiffness is the rough inverse of
compliance. The magnitude of Kn proves to be a bit less
than, but quite close to 1/ZN in good quality unweath-
15.21 ered hard rock, while in the shearing direction,
Ks  1/ZT, sometimes 1/ZT. However ZT data from
Involvement of ZN in the slow shear wave velocity theo- geophysics investigations is extremely limited compared
retically requires dipping fractures, or non-vertical wave to the large body of Ks (pseudo-static) data that has been
propagation. When polarized shear-waves sense the dif- in use in discontinuum rock mechanics modelling since
ferent viscosity of oil or gas in the fractures, ZN is more the late 1960’s, first in jointed FEM studies by Goodman.
likely to be involved than ZT, which would have a less If pseudo-static stiffnesses and dynamic compliances
obvious dependence on fluid viscosity differences. This could be related, despite the different orders of magnitude
sensitivity is despite Gassmann’s theory for porous of dynamic and ‘static’ deformations, then the more
606 Conclusions

researched, stress-and-scale-dependent parameter Ks and 15.27


the stress-dependent Kn might provide shear wave split-
ting analysts more information than at present. A handful The typical 1012 Pa/m unit for stiffness favoured by geo-
of 40 to 50 mm size laboratory samples of natural joints physicists, converts to 1000 MPa/mm, which is more
and stress-induced fractures, and tests on Perspex plates, familiar in rock mechanics. An even better experimental
represent a very uncertain basis for interpreting the equal- feel for stiffness is given by 1 MPa/
m, each of these
ity or inequality ZN  ZT. being equivalent. In geophysics, the inverse of stiffness,
termed compliance is also typically reported in ‘complex’
units such as 1012 m.Pa1, instead of equivalent units
15.24 of 1
m/MPa, which is much more tangible for any rock
mechanics experimentalist.
The measured range of Kn(dyn) / Ks(dyn) ratios for approxi-
mately 50 mm diameter jointed samples in hard crys-
talline rock has been shown from limited testing to be 15.28
about 1 to 4. This means that the ratio of the inverted
compliances ZN/ZT for such samples may range from 1 to Depending on stress levels, rock type, and joint rough-
0.25. Convenient ‘core-sized’ 50 mm data from state-of- ness, Kn(dyn) may range from 1012 to 1014 Pa/m, or 1,000
the art testing of joints in hard crystalline rock seem to 100,000 MPa/mm or 1 to 100 MPa/
m, suggesting a
unlikely to have 1:1 relevance to in situ reservoirs in much mostly very small increment of dynamic displacement.
weaker sedimentary rock, where jointed block sizes may Kn(static) values from a wide range of weaker rock and joint
range from extremes of perhaps 100 mm to 10,000 mm. types, may vary from as low as 100 MPa/mm to almost
50,000 MPa/mm. There is therefore a large degree of
overlap between the static and dynamic stiffnesses in this
15.25 stiffest perpendicular loading direction.

Pyrak-Nolte demonstrated at laboratory scale, that joints


that support less flow tend to have higher normal stiff- 15.29
ness, and that in principal, joint normal stiffness may be
inversely related to the cube root of the permeability. The Shear compliance, as opposed to normal compliance is of
permeability and seismic response of a joint can there- most relevance in the case of vertical shear-waves and ver-
fore be inter-related through the normal stiffness of the tical structure, and the resulting degree of shear-wave
joints in question. Joints that attenuate seismic waves anisotropy. Of the two dynamic compliances, this is the
most at a given stress level, due to lower normal stiff- least understood component. By comparison, it is the
ness, also supported more flow. This implies a second pseudo-static shear stiffness that is most researched in rock
implicit link between the permeability, and both the mechanics. In pseudo-static testing, a whole range of pos-
joint index parameters JRC and JCS, since these can be sible static shear stiffnesses are found, that seem generally to
used to calculate both the normal and shear stiffness, and be inversely related with the sample size, if there is measur-
they independently give estimates of average hydraulic able joint roughness, (and therefore possible permeability
and physical apertures. even at depth). The degree to which dynamic shear com-
pliances might be related, more weakly but nevertheless
directly to sample size, is one of the remaining unsolved
15.26 areas in this important area of seismic detection of
anisotropy, and the subsequent goal of interpreting frac-
The permeability of the rock mass may also be related to tured reservoir permeability.
the rock mass modulus of deformation, when permeabil-
ity and modulus are not reduced by clay. In simplest possi-
ble terms permeability is inversely related to Qrock. Seismic 15.30
quality Qseis and the pseudo-static modulus of deformation
Emass expressed in GPa, can each be estimated from Qrock, Some pieces of the dynamic-permeability jigsaw are com-
for the case of jointed rock masses. plete, but there are missing links between dynamic and
Conclusions 607

static testing, and between small sample testing and the Ks  1 MPa/mm. This low, back-calculated ‘in situ’ shear
large sample reality, which causes experimental stress- stiffness is similar to the values used in large-scale pseudo-
magnitude problems. The completed parts of the jigsaw static modelling of compaction/subsidence in rock
are the abilities to estimate both of the pseudo-static stiff- mechanics. In general one may assume that the pseudo-
nesses Kn and Ks and the less tangible physical (E) and static values of stiffness are lower in the normal direction,
hydraulic (e) apertures, for different sizes of jointed rock and much lower in the shear direction, than the equiva-
block, based on simple index testing. This involves esti- lent dynamic values of stiffness.
mation of joint roughness and wall strength, using respec-
tively the un-scaled or scaled JRC and JCS components of
the Barton-Bandis joint constitutive model. 15.33

Stronger shear wave splitting interpreted from large time


delay in locations in the neighbourhood of fault zones,
15.31
such as the San Andreas fault at Parkfield, suggests that
fluid-filled fractures within the fault zone may be more
Fracture-induced seismic anisotropy has rapidly evolved
extensive than in the surrounding crust. Fault parallel
from the earlier estimation of fracture orientation, with an
polarization of the leading split shear wave may indicate
assumed indication of major horizontal stress, to fracture
that fault-related fractures are aligned by fault shearing
intensity, and the attempted prediction of fluid type, fluid
rather than by the differently aligned regional principal
saturation, and permeability anisotropy. To make this
stress that can be verified further from the fault plane. The
advance, the sensitivity of the fracture compliances to flu-
internal structure of the fault gouge and transition zone is
ids has to be understood. Theoretical expressions for the
assumed to be the reason for this rotation. In some cases,
fracture compliances ZN and ZT have been developed by
only one of the anisotropic shear-wave polarizations is
Liu and Hudson and co-authors, which indicate strong
recorded, due to attenuation of the slower component. A
sensitivity of the ratio ZN/ZT to the bulk modulus of the
change of stress affecting the geometry and fluid in neigh-
fracture infill material, with the most rapid change in the
bouring discontinuities following larger earthquakes, may
compliance ratio, and values closer to 1.0 occurring when
be the reason for temporal decreases in time delay.
the infill bulk modulus approaches zero, such as for gas-
filled fractures. One of the geophysics equations suggests
that ZN/ZT  1 if fractures are dry, and ZN/ZT  0 if frac-
15.34
tures are filled with liquid. With realistically small fracture
aspect ratios (i.e. 0.0001), much lower ratios of ZN/ZT
Earthquakes with shallower focal depths may show pro-
than 1.0 are predicted, which would be more in line with
nounced increase in time delays, suggesting that a
rock mechanics pseudo-static experience. Zero is inadmis-
stronger anisotropy associated with the fault zone may
sible for the pseudo-static ratio Ks/Kn, but certainly
also be concentrated at shallower depth. Leading shear-
Kn Ks.
wave polarizations exhibiting fault-parallel alignments
near the fault, but alignments with the regional stress field
away from the immediate fault zone were also verified at
15.32 the Cajon Pass site. In general terms the source of shear-
wave splitting may be shallower than desired, where con-
Although outside this chapter’s focus on shear-waves, ditions of anisotropy are more favourable for this
Worthington and Hudson modelled the effects on down- mechanism.
going P-waves, of one or more faults intersecting the
transmission path, between 1000 and 2000 m depth in a
North Sea reservoir overburden. They used a theoretical 15.35
compliance model to demonstrate the need for a very
large inequality of the shear and normal compliances, sug- Stress-monitoring sites in Iceland, using state-of-the-art
gesting the need for ZN  4.4  1014 m.Pa1, borehole instrumentation to monitor shear-wave splitting
and ZT  1.1  109 m. Pa1. These convert to shear between controlled-source wells and receiver wells are
and normal stiffnesses of Kn  20,000 MPa/mm, and designed to identify the effects of nearly negligible
608 Conclusions

changes of stress, which might be capable of monitoring of Vp/Vs was 2.9. There was evidence for only a shallow
the build-up of stress and other crustal adjustments before concentration of P-wave anisotropy, which decreased
earthquakes and volcanic eruptions. Anomalous water from 4% at 500 m depth, to zero below 1.5 km depth.
level changes and GPS-measured displacement anomalies Isolated fluid-filled cracks at depths from 500 m to 3 km
have been related to various changes in prior and subse- were too tight to be detected by the P-wave survey, but
quent P-wave and S-wave travel times and delays. could contribute to the shear-wave delays. P-wave
anisotropy was defined as 100  (Vp max  Vp min)/
Vp average, while S-wave anisotropy was defined as
15.36 100  (Vs max  Vs min)/Vs max) following convention.

The ‘90°-flips’ that sometimes occur in source zones, may


be due to extreme pore pressure build-up according to
15.39
Crampin. An alternative explanation is that they could be
due to extreme loading in the H max direction, causing an
Offshore 3D seismic surveys, using compressional P-wave
extreme ‘Poisson effect’ due to crack expansion, thereby
sources converted to PS at the sea-floor, thereafter
locally reversing the H max and H min directions. Except
analysed as S-waves, or use of direct S-waves (SS) gener-
in an eventual clay core, a pore pressure reduction appears
ated on land, in walk-away, multi-azimuth VSP, are basic
more likely than the assumed increase of pore pressure,
geometries for shear-wave polarization and anisotropy
since as shear failure of either intact or jointed rock
investigations above fractured reservoirs. Split shear waves
is approached, dilation is the most likely phenomenon,
S1 (or qS1) and S2 (or qS2 ) may occur due to major prin-
unless beyond the brittle-ductile transition.
cipal stress aligned dominant fracturing: the conventional
interpretation, or perhaps due to stress bisected sets of
unequal conjugate fractures, or due to fracture sets that
15.37
have suffered some permeability-enhancing pre-peak
shear displacement, which may also cause a deviation
Major earthquake (Chi-Chi) after-shock monitoring in
from H max. In walk-around VSP, a circular path of mul-
Taiwan indicated 8% shear-wave velocity anisotropy at a
tiple sources at fixed offset, an incident P-wave, when con-
200 m deep station due to up-going split shear-waves with
verted to a PS wave that passes through the plane of
fast and slow components. Qseis values in the 2–15 Hz fre-
horizontal symmetry caused by aligned vertical fractures,
quency band were 61 to 68 in the fast direction, and
will show polarity reversal when incident on either side of
43–52 in the slow direction. Such values may closely
the fracture strike.
resemble possible deformation moduli M, expressed in
GPa. The anisotropy may have been caused by sub-verti-
cal joint set anisotropy, which causes anisotropy at low
seismic frequencies, as opposed to the assumed micro- 15.40
crack anisotropy that is dominant from 1 kHz.
The overburden may also display anisotropy in the form
of azimuthal-dependent velocity. Errors will be intro-
15.38 duced when the seismic data are inverted to obtain frac-
tured reservoir parameters, if overburden anisotropy is
Shear-wave splitting analysis utilizing local 4 to 5 km deep ignored. Shear-wave splitting and polarization occurs in
micro-earthquakes, and P-wave anisotropy measurement both cases due to the presence of the relatively compliant
with controlled sources, were used in Mid-Atlantic Ridge fracture properties. For the reservoir, basic structural
studies, where anisotropy was attributed to a shallow dis- information such as fracture density, strike and dip (due to
tribution of vertical, fluid-filled cracks, aligned parallel to symmetries), and some indications of fluid-type (gas or
the trend of the axial valley. Time delays ranged from 35 oil) and permeability may also be obtained, due to the rel-
to 180 ms, with an average delay of 90 ms. The shear-wave ative sensitivities of the fracture normal and shear compli-
anisotropy was from 8 to 30% in the highly fissured layer. ances. Deviation of fracturing from H max, because one of
Most of the shear wave delay was attributed to the shal- the sets dominates, or because fracture set stiffnesses differ,
lowest 500 m in seismic layer 2A, where the average value may compromise relative crack density interpretation.
Conclusions 609

15.41 high. The response is strongest when fractures are gas


filled, as gas does not stiffen the normal compliance.
Conjugate fracturing with non-vertical dips, as typically Attenuation anisotropy should be enhanced in the case of
found in anticlinal or domal structures, means that sam- oil-filled fractures, due to the greater contrast of their nor-
ples of both the oppositely-dipping joint sets can be mal and shear compliances.
directly sampled, and tested, using vertical exploration
wells. Nevertheless, it is a remarkable fact that exploration
wells are nearly always vertical despite the target structures
15.44
often being vertical too. Sidewall cores are therefore used
to help locate zones having high fracture intensity; using
Reportedly the world’s first time-lapse, marine, multi-
microfracture and diagenesis data to infer the presence of
component (3D/4C) survey, was performed in 2002 at
the macro-fractures.
the Ekofisk jointed-chalk reservoir in the North Sea. This
baseline was subsequently compared with a monitoring
4D/4C survey acquired in 2003. Small changes of polar-
15.42 ization direction and of shear-wave anisotropy were
detected, even after this short 15 months monitoring of
Large aperture fractures may require mineral bridges to more than 30 years production. Azimuthal rotation of the
have preserved conducting porosity. Fractures below a anisotropic attenuation has recently also been detected at
certain characteristic size may be completely filled. ‘Open’ neighbouring Valhall, where life time seismic monitoring
fractures in the sub-surface are not necessarily parallel to is established. An unequal conjugate shear mechanism
maximum compressive stress H max. Fractures perpendi- could be a possible explanation for these temporal rota-
cular to this direction can also be ‘open’ if partially filled tions, especially if the strike of the two conjugate sets was
with synkinematic or post-kinematic mineral cements. not equally oriented. Potential ‘opposite-rotation’ of fluid
Sealed fractures parallel to H max may also be numerous. lenses and rock-to-rock contact areas due to pre-peak
Effective normal stress induced joint ‘closure’ caused by shearing of non-planar joints might also be detected by
sufficiently high h min may occur at reservoir depths in the shear waves. A much larger rotation of polarization
less competent rock. There is however an important pre- azimuth was detected at a monitored HDR in Cornwall,
peak shear-displacement mechanism for maintenance of where joint shearing was also assumed, due to plunging
‘open’ fractures that are non-parallel to H max, in which AE activity.
conjugate sets may be bisected by H max. This interpreta-
tion has numerous merits for explaining permeability at
depth.
15.45

An important development in the dynamic modelling


15.43 of the effects if fractures is that of Schoenberg, who
modelled elastic wave behaviour using linear slip inter-
When fracture densities are as high as 1 to 2, as in well- faces. These allow reflection, transmission, conversion,
jointed, domal chalk reservoirs, dimming of the ampli- and delay to take place at the modelled interface, with
tudes of the slow shear-wave, due to greater attenuation the magnitudes depending on the specific stiffness, the
caused by lower Qseis tend to correlate with the most pro- frequency content, and the angle of incidence. The
ductive parts of the reservoir, as also experienced where assumption is that when an elastic wave propagates
the measurable shear-wave anisotropy is greatest. A across a fracture, there is a displacement discontinuity
down-dip shear mechanism may help to maintain aper- that is linearly related to the normal or shear force gen-
tures despite high effective stresses in the presence of this erated. The seismic particle displacement is discontinu-
weaker, high porosity rock. If the high crack density is ous, while the seismic stresses are assumed to be
contributed to by two sets of oppositely dipping conju- continuous. This shares several of the interface-stiffness
gate joints, one can expect shear and normal compliance concepts that are basic to discontinuum modelling in
contributions from both sets to the slowness of the slow rock mechanics, in the Cundall 2D and 3D codes
shear-wave, which might be 2 km/s when porosity is UDEC and 3DEC.
610 Conclusions

15.46 fluid transfer between the different scales of discontinuity,


dispersion occurs at lower frequency ranges than those
Hudson utilised a ‘method of smoothing’ for the effect over which the micro-structure dispersion occurs.
of modelled cracks, which was capable of representing Introduction of even small fractures causes dispersion to
the elastic parameters of a ‘cracked’ material, in the form begin at lower frequencies, and the effect increases with
of an effective medium. This allowed calculation of the increasing fracture sizes. The magnitude of shear wave
effect of incident dynamic waves of long wavelength. anisotropy therefore shows fracture size, frequency and
Subsequently, Hudson extended the model to allow azimuth dependence. Attenuation is altered by setting
for the cracks to be connected via the porosity of a rock smaller aspect ratios for meso-fractures compared to
matrix. In this extended case, cracks could be deformed microcracks, as a result of higher stress, as the fluid cannot
by an incident wave in a manner that depended on their ‘squirt’ or flow so easily, in response to the passage of seis-
aspect ratio and on their orientation with respect to the mic waves.
incident wave. Clearly the modelling of intrinsic attenua-
tion mechanisms such as squirt flow, and its frequency
dependence, was transformed by this extended capability. 15.49

Effective medium models appear at present to make no


15.47
distinction between the assumed mean crack aperture (E),
and the theoretical hydraulic aperture (e). In jointed rock
Tod extended this model further, by allowing for a con-
(E) controls stiffness and deformation moduli, and the
tinuous distribution of crack orientation and aspect ratios,
crack aspect ratio should have a similar role. On the other
and by allowing each to depend on the applied stress
hand, (e) controls the intrinsic permeability, given as
and on the fluid pressure. At high frequencies, the
e2/12. Both apertures probably influence attenuation,
cracks behave as isolated stiff features, while squirt
since they effectively define two different aspect ratios,
losses at lower frequencies cause a soft behaviour. Crack
therefore influencing the assumed squirt-flow losses and
density is allowed to decay with an increase in applied
the assumed stiffness. Dispersion will likely begin at
stress, from an initial value representing the unstressed
higher frequencies as roughness increases. Incorporating
state. The model is purely elastic, and relaxes to its orig-
the inequality e  E would be a source of improvement
inal state upon unloading. Crack density is designed to
for effective medium modelling of rough-walled, tightly
decrease with increasing compressive stress, but cracks
compressed cracks, joints or fractures, for which use of
with normals lying in the h minimum direction are
only one aspect ratio is least realistic.
assumed to remain open. The model is capable of cap-
turing the changed anisotropy caused by fluid pressures
and applied stresses, which impact the aspect ratios of
15.50
the cracks. The non-compliant pores of the Tod model,
transfer fluid to the physically unconnected cracks, there-
The traditional EDA focus on microcracks as the source
fore giving dispersive, or frequency-dependent velocities
of shear-wave splitting, may need to be re-evaluated, in
and attenuation.
the light of these new poro-elastic double (or triple) poros-
ity models. Microcracks appear to give a ‘constant’
15.48 potential source of shear-wave splitting independent of
frequency. Meso-fractures, for instance of 1 m and 10 m
Chapman developed a triple-porosity poro-elastic model, radius, apparently have an equally strong role in shear-
based on the observation that typical laboratory samples, wave anisotropy as microcracks, but specifically at the
clearly unfractured, nevertheless display dispersion, seismic frequencies appropriate to earthquake studies.
anisotropy, stress sensitivity, and dependence on fluid type
and degree of saturation. When the fractures or cracks
were removed from preceding models, a linear-elastic 15.51
material remained, in contrary to observation. Chapman
combined ‘meso-scale’ anisotropic fractures with equant Shear wave splitting through fractured reservoirs may
matrix porosity and ellipsoidal microcracks. Due to the show a sharp increase of time delay with depth at the
Conclusions 611

reservoir level, indicating the presence of fractures. The geomechanics concept is provided by deep-well data
time delays tend to be largest at lowest frequencies (e.g. showing that joints under differential (shear) stress are
5–15 Hz), and smallest at higher frequencies (e.g. the conductors, with other directions apparently non-
20–40 Hz). However, polarization of the fast S-waves may conducting.
show no apparent variation with frequency, unless at very
low frequency. The time delays between the split shear
waves may be decreasing as frequency increases, due to 16.3
stiffening of the squirt phenomenon in the case of the
slow waves. Such data have been used for inverting for the The interpretation of shear wave splitting with possible
theoretical fracture density (an appropriate re-naming of multiple ‘vertical’ joint or fracture orientations emphasises
crack density), and for fracture radius, using nine-compo- the need for knowledge of both shear and normal compli-
nent VSP data sets, with one P-wave and two orthogonal ances, if poro-elastic modelling is to extend to stress and
S-wave sources. Frequency-dependent anisotropy has also displacement sensitive aspect ratios for the modelled joints
been demonstrated. or fractures. Presently, limited sets of laboratory data, and
even more limited sonic log and cross-hole data, and a
tentative extrapolation to large scale seismic interpreta-
Chapter 16 Joint stiffness and tion of fault compliances, strongly suggest a scale effect.
compliance and the joint Increased compliance in both normal and shear directions
shearing mechanism is implied, as scale is increased or frequency reduced.

16.1 16.4
Rock mechanics developments in distinct element rock A comparison has been made between the presently
mass deformation modelling, including flow within the known normal ZN and shear ZT compliance laboratory
deforming joints, can be used to illustrate an extension of and field data, which is a dynamic and micro-deforma-
the traditional geophysics concept of one set of stress- tion response, and the much larger body of laboratory
parallel ‘open’ joints in a reservoir. This is important and field data for the pseudo-static response of normal
because multiple sets of joints are more usual in rock Kn and shear Ks stiffness. Data for joints, tension frac-
masses, probably including even deep reservoirs. Multiple tures and clay-filled discontinuities are readily available
joint or fracture sets, such as bedding and two vertical sets due to long term use in rock mechanics modelling of
also open the possibility of polarization orientations that jointed rock masses. These pseudo-static measures of
are not parallel to the H max direction. Two conjugate stiffness are given in typical units of MPa/mm, and when
vertical joint sets intersected by the H max direction can inverted show resemblance but generally larger magni-
also cause shear wave splitting with polarization that tude, compared to the geophysicists compliance, given in
nearly corresponds to this principal stress direction. typical m/Pa units.

16.2 16.5

When reservoir joints or fractures are not parallel to the When rock mass and rock joint quality is high, the
H max direction, they may be acted on by shear or differ- dynamic modulus Edyn is not of much greater magni-
ential stress, and this implies shear deformation, which tude than the static modulus Emass. The normal compli-
may be needed to supplement the often too small con- ance ZN may then be within 1 to 1/10 of the magnitude
ducting apertures which may result when in the tradi- of the inverted static normal stiffness Kn. A typical lab-
tional parallel to H max direction. This philosophy is based oratory-scale ZN range of 1013 to 1014 m/Pa, or Kn
on the fact that unless joints or fractures are mineral (dyn) of 10,000 to 100,000 MPa/mm is therefore found
‘bridged-but-not-blocked’, or are very rough and in hard to be 1 to 10 times stiffer than typical Kn (static) data
rock, both testing and modelling indicates too small aper- of typically 1,000 to 10,000 MPa/mm (range maybe
tures to be considered ‘open’, since conducting apertures 100 to 50,000 MPa/mm). In the inverted worlds of
may be 5
m, and often 1
m. Support for this geophysics and rock mechanics, 1012 m/Pa is the same
612 Conclusions

as 1,000 MPa/mm, with the identical 1 MPa/


m giving stress reduction from a higher stress level, if sufficiently
a more practical feeling for this level of micro dynamic rough and in sufficiently strong rock. Cooling can have a
stiffness. similar effect. Both ambient and thermal over-closure
cause higher shear strength to be registered, due to
changes in the thermal expansion coefficient, when joints
16.6 are included. This causes changes in aperture and perme-
ability, and may therefore complicate the matching of
Although the large-scale, seismically derived magnitude reality and present theory. During heating, stiffness and
of shear compliance ZT from reservoirs is presently very modulus appear to be less than expected, due to the addi-
uncertain, preliminary indications are that it may be tional closure. However, following the heating, this ther-
about two orders of magnitude stiffer than large-scale mal over-closure results in stiffer subsequent behaviour.
pseudo-static shear stiffness Ks. The lower magnitude of
Ks is due to strongly reducing magnitudes as block size
increases. A block size increase has a double effect due 16.10
to reduced peak shear strength and larger displacement
to reach peak strength. It is not known whether this fun- Despite insufficient test data from most areas, the utiliza-
damental scale effect could also influence the presumed tion of index properties for the joints for guiding the esti-
stiffer dynamic shear stiffness, but a weaker dependence mation of Kn and Ks, which is well established in rock
than in rock mechanics is suspected. mechanics, could provide order of magnitude estimates of
the dynamic properties, or estimates of the index proper-
ties and therefore permeability, by inversion. This would
16.7 presently be based on the assumed one to two order of
magnitude stiffer dynamic behaviour. Index properties
A scale effect on ZN or Kn (dyn), as occurs with Ks, is not JCS and JRC concerning wall strength and roughness,
expected when merely sampling a larger portion of the allow estimation of the mechanical and hydraulic aper-
same feature, However, when comparing ZN or Kn (dyn) tures under closure and shearing, and also estimation of
on small fractures, larger joints, or faults, a sampling effect the pseudo-static normal and shear stiffnesses each,
must obviously be expected, since the inequality thereby potentially linking dynamic data and in situ
Edyn  Emass is potentially increasing when sampling permeability.
increasingly large features.

16.11
16.8
Index data for typical reservoir rocks and joints, with vari-
There is present uncertainty in the magnitudes of ZN and ation of both strength JCS and roughness JRC, and con-
ZT because of limited test data at different scales. The sideration of confinement effects on strength, give
uncertainty in the magnitudes of the pseudo-static Kn prediction of very small conducting apertures of micron
and Ks parameters used in rock mechanics is much less, or less magnitude when modelling closure by an assumed
and values can be readily predicted using a 1D model for h min of 10 to 40 MPa, unless roughness and strength are
coupled M–H joint behaviour. However, there is an both significant. Available coupled M–H test data from
inevitable lack of test data with respect to high pressure the laboratory CSFT apparatus, and from in situ block
and large size, also in the case of pseudo-static testing. tests, each including heating, also confirm the extremely
small apertures of interlocked joints, unless rough and of
high strength, or with aperture preserved by mineral
16.9 bridging.
A further uncertainty with all four joint parameters is that
rough joints or fractures tend to be more tightly closed at 16.12
high temperature, and are also more tightly closed follow-
ing several load cycles if they have been sampled and When modelling minor amounts of shear, excellent per-
unloaded, and cooled. Joints can appear ‘over-closed’ by meability is predicted in most cases, due to the positive
Conclusions 613

influence of even pre-peak-strength shear displacements, 16.15


showing the vitally important beginnings of dilation. One
to two orders of magnitude increased permeability is seen Using non-linear Barton-Bandis shear strength, dilation
in CSFT coupled shear testing, following only 1 or 2 mm and aperture interpretation, in place of the linear Byerlee
of pre-peak or close-to-peak shear displacement. Shearing
 0.6 to 1.0 assumption for the shear strength of
of non-planar surfaces causes the fluid-bearing lenses on ‘faulting’, one finds that the conducting ‘fractures’ with

‘down-slopes’ to have a different average orientation than in the range 0.6 to 1.0, require typical to high ranges of
the ‘up-slope’ rock-to-rock contact areas. This rotation both JRCn, and confined strength JCSn. This means a
may influence the polarization direction of split shear full-scale roughness as high as 10, and full scale confined
waves, if they are more sensitive to one or the other of rock strengths ( 1  3) as high as 200 to 600MPa or
these average directions. Combined with two conjugate more, to explain this higher range of resisted / n loading.
but usually unequally developed joint or fracture sets, the The subscripts (n) on the index parameters signify in situ
resultant polarization directions of the split shear waves strength, and block sizes Ln of perhaps 0.25 to 2.5m.
qS1 and qS2, can have at least two reasons for not lining up
exactly with the classically ‘preferred’ H max and h min
directions, and for showing 4D rotations of azimuth. 16.16

It is very difficult to model


as high as 1.0 in situ, in
jointed or fractured rock at kilometre depths, but easy
16.13
when near-surface. The joints or fractures must be pre-
peak or close to peak strength to have sufficient strength
Significant contributions from deep-well monitoring and
to develop implied mobilized friction angles as high as
analyses by Colleen Barton, Zoback and Townend and
45°. At the same time they must be sufficiently dilated
colleagues, has demonstrated the hydraulic flow distinc-
to be conducting. ‘Fractures’ that are under the lower

tion of joints or fractures that are under differential shear


range of 0.4 to 0.6 are more likely to have the character
stress, and those that are principally under larger normal
of minor faults or have larger block sizes or some clay
stress and insignificant shear stress, due to their mutually
smear.
different orientations. In general, but quite clearly, the
former are found to be water conducting, and the latter
are assumed not to be water conducting, based on tem-
16.17
perature logging. The wells are of kilometre to several
kilometre depths, drilled in mostly crystalline rocks, with
The assumption of co-axial stress and displacement in
several wells connected with San Andreas Fault investiga-
the classic stress-transformation equations seems to be in
tions. This hydraulically differentiated behaviour is
error if the joints or fractures or new fault surfaces are
despite rock strengths significantly higher than typical
non-planar. Since shear displacement is implied when
reservoir rocks.
specifying shear strength, or a resisted value of / n 
,
then dilation has also inevitably occurred at the highest
range of
from about 0.7 to 1.0, for which significant
16.14 roughness or non-planarity is also implied.

Differential stress magnitudes based on rock stress meas-


urements, indicate that the presently resolved ratios of 16.18
shear and effective normal stress expressed as mobilized
frictional strength
, are generally in the range of 0.4 to The mobilized dilation angle dn mob estimated from the
1.0 for the case of the above water conducting fractures. JRCmobilized concept, needs to be added to the angle 
On this basis fractured reservoir jointing that was paral- used to define the joint or fracture orientation in relation
lel to H max and without mineral ‘bridging’ or sufficient to the major principal stress direction 1. The addition of
JCS and JRC might well be very impermeable when the dilation as sin 2( dn mob) and cos 2( dn mob) in
normally pressured, but perhaps conducting when over- the shear and normal stress transformation equations
pressured. explains the extra difficulty of shearing when dilation
614 Conclusions

occurs, and might explain the high end of the interpreted 16.20
values of in situ
.
Due to the problem of clay-sealing of discontinuities,
and due also to the general effect of reduced permeabil-
16.19 ity at significantly increased effective normal stress, a new
term called QH20 has been developed, involving an
An approximate, order of magnitude prediction of per- inverted Jr/Ja term, a normalized JCS, together with a
meability in rock masses, can be made using the rock simple depth-permeability equation for the ‘soft poros-
mass Q-value, normalized by the uniaxial strength to ity’ represented by jointed rock. This shows promising fit
the form Qc. This can be equated to the inverse of the to shallow civil engineering Lugeon testing, and poten-
Lugeon value, where 1 L is 107 m/s. The resulting tially also to deep-well data, and demonstrates suitable
Qc  10, K  108 m/s type of estimate appears to be adjustment to lower permeabilities caused by clay-filled
a useful first order estimate, when Qc values range from discontinuities and increased depths.
0.1 to 1000, implying little complication of clay smear.
The range of K is then predicted to be about 106 to
1010 m/s, or roughly 1013 to 1017 m2 if water vis-
cosity at 20°C is invoked, for simplicity.
Appendix A – The Qrock parameter
ratings

RQD J J Definitions of characterization


Q   r  w and classification as used in
Jn Ja SRF rock engineering

CHARACTERIZATION description of a virgin site


The six parameters defined (pre-tunnelling attributes
and properties)
RQD is the % of competent drill-core sticks CLASSIFICATION description of a non-virgin
100 mm in length in a selected domain site (post-excavation attrib-
(Deere et al., 1967) utes and properties)
Jn  the rating for the number of joint sets (9 for 3 For example, in the excavation disturbed zone of a tunnel
sets, 4 for 2 sets etc.) in the same domain or steep rock slope, there will be changes in stress, per-
Jr  the rating for the roughness of the least fav- meability, deformation modulus and seismic velocity. In
ourable of these joint sets or filled disconti- the case of a tunnel, it may not only be the redistributed
nuities stresses that have radial and tangential components. With
Ja  the rating for the degree of alteration or clay rock joints present, four sectors of shear stress and joint
filling of the least favourable joint set or filled displacement (roughly at 45° intervals) and the pairs of
discontinuity diametrically opposite maximum and minimum tan-
Jw  the rating for the water inflow and pressure gential stress, may give a complex perturbation of prop-
effects, which may cause outwash of discon- erties in the EDZ.
tinuity infillings The tables appearing in the following figures are the
SRF  the rating for faulting, for strength/stress ratings for the six ‘Qrock’ parameters. These have been
ratios in hard massive rocks, for squeezing or printed as figures, in order to keep the compact style suit-
for swelling, as appropriate able for reproducing together with field logging sheets.
The recommended way of recording the Q-parameter
ratings is explained in the following notes, based on
Combination in pairs Barton, 2002a. Footnotes below the tables that follow,
also give advice for site characterization ratings for the case
RQD / Jn  relative block size (useful for distinguish- of Jw and SRF, which must not be set to 1.0 and 1.0, as
ing massive, rock-burst-prone rock) some authors have suggested. This destroys the intended
Jr / Ja  relative frictional strength (of the least multi-purposes of the Q-system, which has a quite differ-
favourable joint set or filled discontinuity) ent structure compared to RMR. (Bieniawski, 1989)
Jw / SRF  relative effects of water, faulting, strength/
stress ratio, squeezing or swelling (‘active
stress’)
Notes on Q-method of rock mass
An alternative combination of these three quotients in classification
two groups only, has been found to give fundamental
properties for describing the shear strength of rock 1. These tables contain all the ratings necessary for
masses – something close to the product of ‘c’ and ‘tan classifying the Q-value of a rock mass. The ratings
␸’. By implication Q (and in particular Qc) have units form the basis for the Q, Qc and Qo estimates of
resembling MPa. (Barton, 2002a) rock mass quality (Qc needing only multiplication
616 Rock quality, seismic velocity, attenuation and anisotropy

of Q by c /100, and Qo the use of a specifically ori- section of tunnel, for separate analysis and reporting.
ented RQD, termed RQDo relevant to a loading or Overall frequencies of observations of each rating
measurement direction). All the classification rat- (or selected sets of data) can be given as numbers on
ings needed for tunnel and cavern design are given separate logging sheets. Large data sets can be plot-
in the six tables, where Q only would usually apply. ted in e.g. EXCEL when returning from the field.
2. For correlation to engineering parameters as described 5. It is convenient and correct to record rock mass
in this paper, use Qc (multiplication of Q by c / variability. Therefore allow as many as five observa-
100). For specific loading or measurement directions tions of each parameter, for instance in a 10 m
in anisotropically jointed rock masses use RQDo in length of tunnel or 5 m length of core. If all obser-
place of RQD in the Q estimate. This means that vations are the same, great uniformity of character
an oriented Qc value should contain a correctly is implied, if variable – this is important informa-
oriented RQDo for better correlation to oriented tion. At ‘the end of the day’ the histograms will give
engineering parameters. a correct record of variability, or otherwise.
3. Q-parameters are most conveniently collected using 6. Remember that logged RQD of 10, including 0,
histogram logging. Besides space for recording the are set to a nominal 10 when calculating Q, to
usual variability of parameters, for structural domain avoid calculating Q  0. In view of the log scale of
1, domain 2 etc., it contains reminders of the tabu- Q, the histograms of RQD in the logging sheet will
lated ratings at the base of each histogram. Space for be sufficiently accurate if given mean values, from
presentation of results for selected (or all) domains left to right, of 10, 15, 25, 35…85, 95, 100. The
at the top of the diagram, includes typical range, log scale of Q also suggests that decimal places
weighted mean and most frequent (Q-parameters, should be used sparingly. The following is consid-
and Q-values). ered realistic 0.004, 0.07, 0.3, 6.7, 27, 240. Never
4. During field logging, allocate running numbers to report that Q  6.73 or similar, since a false sense
the structural domains, or core boxes, or tunnel of accuracy will be given.
sections, e.g. 1  D1, 2  D2 etc. and write the 7. Footnotes below the tables that follow, also give
same numbers in the allotted histogram columns, advice for site characterization ratings for the case
using a regular spacing for each observation such as of Jw and SRF, which must not be set to 1.0 and
11, 113, 2245, 6689 etc. In this way the histograms 1.0, as some authors have suggested. This destroys
will give roughly the correct visual frequency of all the intended multi-purposes of the Q-system, which
the assembled observations, in each histogram col- has an entirely different structure compared to RMR
umn. Besides this, it will be easy to find the relevant (Leave blank if permeability and stress data is
Q-parameters for a particular domain, core box or awaited, otherwise estimate Jw and SRF.)
Appendix A – The Qrock parameter ratings 617

Figure A1 Characterization/classification ratings for RQD, Jn and Jr.


618 Rock quality, seismic velocity, attenuation and anisotropy

Figure A2 Characterization/classification ratings for Ja and Jw.


Appendix A – The Qrock parameter ratings 619

Figure A3 Characterization/classification ratings for SRF.


620 Rock quality, seismic velocity, attenuation and anisotropy

Figure A4 Graphic presentation of the meaning of Jr/Ja, representing the frictional strength of joints and clay-filled discontinuities. Note
tendency for friction angle development like   i, , and  – i, according to whether dilatant or ‘normal’, or contractant joint
or discontinuity resistance to shearing.
Appendix A – The Qrock parameter ratings 621

Figure A5 The Q-system of tunnel (and rock cavern) permanent support estimation, based on Grimstad and Barton, 1993, and Barton,
2000. The other widely used ‘rock mass rating’ (RMR) used in engineering geology has approximate correlation to Qrock as shown
in the equations. The version ‘RMR  15 log Q  50’ is preferred. Barton, 1995. (Note: B  systematic bolting i.e. grouted
steel rebar. Sfr  fibre reinforced sprayed concrete, often called shotcrete).
622 Rock quality, seismic velocity, attenuation and anisotropy

Figure A6 The Q-parameter histogram logging sheet, for recording the number of observations of each parameter.
Appendix A – The Qrock parameter ratings 623

Figure A7 Example of hand-filled Q-parameter ratings from core-logging of part of a deep borehole.
624 Rock quality, seismic velocity, attenuation and anisotropy

Figure A8 Example of EXCEL calculation of rock mass quality statistics, for the deep borehole, part of which was logged in Figure A7.
Appendix B – A worked example
Q-value and tunnel or cavern support 2. Emass (or M)  10 GPa, Vp (seismic)  3.5 km/s,
Qseis  12. (A rock mass porosity equal to a nomi-
Note the overbreak caused by three joint sets in Figure B1. nal 1% has been assumed).
The joint planes (beneath the sprayed concrete) are planar 3. From Figures 5.36 and 5.37, the empirically-derived
and weathered or clay-coated, making for poor stability effect of increased depth can be traced; i.e.
until also reinforced with rock bolts. Q  90/9  Vp  5 km/s, and Emass  32 GPa at 500 m depth.
1/4  0.66/2.5  0.7 (‘very poor’). Estimate of perma- 4. By implication, with a (continued, pessimistic)
nent support requirements from Figure A5: 20 m span assumption of unchanged rock mass quality with
requires B (bolting) 1.6 m c/c (spacing)  13 cm S(fr) increased depth, the magnitude of Qseis (specifi-
(steel-fibre reinforced ‘shotcrete’ or sprayed concrete). cally Qp at seismic frequency) would be expected to
be about 30.
Q-value used for geophysical 5. The reality of improved Q-value at depth (e.g.
estimates Jr/Ja  2/1, and SRF  0.5 (high stress) would
mean Q  40. With less well-developed joint sets,
1. With a simple c estimate for the gneiss of 150 MPa, the rock mass quality Q-value could easily be 100,
Qrock  0.66  150/100  1.0. Therefore at this at 500 m depth.
shallow (25 m deep) cavern site the following ‘geo- 6. With the 40–100 estimate of Qrock, a more likely
physical’ estimates can be made, from Figure 5.36, range of Vp is about 5.6–6.0 km/s, with Qseis
5.37, 13.60 and 15.33. increased to about 60–68 at 500 m depth.

Figure B1 The reality of near-surface construction of tunnels and caverns in rock. Note the three joint sets causing deep over-break. See Plate 17.
626 Rock quality, seismic velocity, attenuation and anisotropy

7. The approximate Qseis estimates of about 10 at ‘explanation’ for this simple (probably too simple) link
25 m depth, and about 60–70 at 500 m depth are is that all three parameters (Qrock and Qseis and Emass)
mirrored numerous times in this book. are reflecting the ‘soft-porosity’ effects of both structure
and joint stiffness, with potentially several joint sets
Qrock and Qseis appear to be approximately linked via involved, not purely normal loading across one set. For
the pseudo-static deformation modulus, which can be some reason, rock mass dynamic stiffness as reflected in
estimated from Qrock, as we have seen. The possible Edyn, is too high for a good correlation.
References

A Allan, J.R. 2002. Controls on reservoir performance during


primary and enhanced recovery: lessons learned from 100
Abercrombie, R.E. 1997. Near-surface attenuation and site fractured giant fields. American Association of Petroleum
effects from comparison of surface and deep borehole Geologists (AAPG Reservoir Geology Committee), Annual
recordings. Bull. Seism. Soc. Am., 87: 731–744. Meeting, Houston: TX.
Abercrombie, R.E. 1998. A summary of attenuation measure- Alt, J.C., Honnerez, J., Laverne, C. & Emmermann, R.
ments from borehole recordings of earthquakes: the 10 Hz 1986. Hydrothermal aleration of a 1-km section through
transition problem. Pure Appl. Geophys., 153: 475–487. the upper oceanic crust. Deep sea drilling project hole
Abercrombie, R.E. 2000. Crustal attenuation and site effects 504B. J. Geophys. Res., 91: 10,309–10,335.
at Parkfield, California. J. Geophys. Res., 105: 6277–6286. Amadei, B. & Stephansson, O. 1997. Rock Stress and its
Adams, D.A. & Abercrombie, R.E. 1998. Seismic atten- Measurement. 490 pp. London, New York: Chapman & Hall.
uation at high frequencies in Southern California from Amaefule, J.O., Altunbay, M., Tiab, D., Kersey, D.G. &
coda waves recorded at a range of depths. J. Geophys. Res., Keelan, D.K. 1993. Enhanced reservoir description: using
103: 24,257–24,270. core and log data to identify hydraulic (flow) units and pre-
Adams, L.H. & Williamson, E.D. 1923. The compressibility dict permeability in uncored intervals/wells. SPE Paper
of minerals and rocks at high pressures. J. Franklin Inst., 26436, 1–16. Tulsa, OK: Society of Petroleum Engineers.
195: 475–529. Anderson, R.N., Zoback, M.D., Hickmann, S.H. &
Addis, M.A., Barton, N., Bandis, S.C. & Henry, J.P. 1990. Newmark, R.L. 1985. Permeability versus depth in the
Laboratory studies on the stability of vertical and deviated upper oceanic crust: in situ measurements in DSDP hole
boreholes. 65th Annual Technical Conference and Exhibition 504B, Eastern Equatorial Pacific. J. Geophys. Res., 90:
of the Society of Petroleum Engineers, New Orleans. 3659–3669.
Agarwal, Y.P., Sykes, L.R., Armbruster, J. & Sbar, M.L. Angerer, E., Crampin, S., Li, X.-Y. & Davies, T.L. 2002.
1973. Premonitory changes in seismic velocities and pre- Processing, modelling and predicting time-lapse effects of
diction of earthquakes. Nature, 241: 101–104. overpressured fluid-injection in a fractured reservoir.
Akbar, N., Dvorkin, J. & Nur, A. 1993. Relating P-wave Geophys. J. Int., 149: 267–280.
attenuation to permeability. Geophysics, 58: 20–29. Aoki, K., Shiogama, Y., Tezuka, Y., Kobuchi, T. &
Akbar, N., Mavko, G., Nur, A. & Dvorkin, J. 1994. Seismic Masumoto, K. 1991. Evaluation of hydraulic conductivities
signatures reservoir transport properties and pore fluid of jointed rock mass by cross-hole permeability test. Int.
distribution. Geophysics, 59: 1222–1236. Congress on Rock Mechanics ISRM, 7. Aachen 1991. 1:
Aki, K. 1969. Analysis of the seismic coda of local earth- 423–426. Rotterdam: Balkema.
quakes as scattered waves. J. Geophys. Res., 74: 615–631. Archambeau, C.B., Flinn, E.A. & Lambert, D.G. 1969. Fine
Aki, K. & Chouet, B. 1975. Origin of coda waves: source, structure of the upper mantle. J. Geophys. Res., 74: 5825–
attenuation, and scattering effects. J. Geophys. Res., 5865.
80(23): 3322–3342. Armstrong, T., McAteer, J. & Connolly, P. 2001. Removal of
Aki, K. & Richards, P.G. 1980. Quantitative Seismology: Theory overburden velocity anomaly effects for depth conversion.
and Methods. 932 pp. San Francisco: W.H. Freeman and Co. Geophys. Prospect., 49: 79–99.
Akinci, A., Mejia, J. & Jemberie, A.L. 2004. Attenuative dis- Arts, R.J., Rasolofosaon, P.N.J. & Zinszner, B.E. 1996.
persion of P-waves and crustal Q in Turkey and Germany. Experimental and theoretical tools for characterizing
Pure Appl. Geophys., 161: 73–91. anisotropy due to mechanical defects in rocks under varying
Albert, W. 2000. Petrophysical parameters – an indication of pore and confining pressures. Fjær, Holt and Rathore (Eds.).
rock permeability? Nagra Bulletin, 26: 26–37. Seismic Anisotropy: 384–432. Society of Exploration
Albright, J., Dangerfield, J., Johnstad, S., Cassell, B., Geophysicists.
Deflandre, J.-P. & Withers, R. 1994. Seismic surveillance for Assefa, S., McCann, C. & Sothcott, J. 1999. Attenuation of P-
monitoring reservoir changes. Oilfield Rev., January 1994, and S-waves in limestones. Geophys. Prospect., 47: 359–392.
Schlumberger. Aydan, Ö., Akagi, T., Ito, T. & Kawamoto, T. 1992. Prediction
Al-Chalabi, M. 1997. Parameter nonuniqueness in velocity of behaviour of tunnels in squeezing ground. J. Geotech.
versus depth functions. Geophysics, 62(3): 970–979. Eng., 448(III-19): 73–82. JSCE.
628 References

B Barkved, O., Bartman, B., Gaiser, J., Van Dok, R., Johns, T.,
Kristiansen, P., Probert, T. & Thompson, M. 2004. The
Bachrack, R. & Nur, A. 1998. High-resolution shallow- many facets of multicomponent seismic data. Oilfield Rev.
seismic experiments in sand, Part I: Water table, fluid flow, Summer, 2004, 42–56, Schlumberger.
and saturation. Geophysics, 63(4): 1225–1233. Barkved, O.I. & Kristiansen, T. 2005. Seismic time lapse
Bachrach, R., Dvorkin, J. & Nur, A. 2000. Seismic velocities effects and stress changes: examples from a compacting
and Poisson’s ratio of shallow unconsolidated sands. Geo- reservoir, The Leading Edge, December 2005, 1244–1248.
physics, 65(2): 559–564. Barla, G. 1993. Case study of rock mechanics in the Masua
Badri, M. & Mooney, H.M. 1987. Q measurements from Mine, Italy. Comprehensive Rock Engineering, Vol. 5.
compressional seismic waves in unconsolidated sediments. Surface and Underground Case Histories. 291–334.
Geophysics, 52(6 ): 772–784. Oxford: Pergamon.
Bakhtar, K. & Barton, N. 1984. Large scale static and dynamic Barton, C.A., Zoback, M.D. & Moos, D. 1995. In situ stress
friction experiments. Rock Mechanics in Productivity and and permeability in fractured and faulted crystalline rock.
Protection, Proc. 25th Symp. Rock Mech., Evanston, IL, Mechanics of Jointed and Faulted Rock, Rossmanith (Ed.).
Dowding, C.H. (Ed.). The American Institute of Mining, 381–396. Rotterdam: Balkema.
Metallurgical and Petroleum Engineers (AIMM). Barton, N. 1971. A model study of the behaviour of steep
Bakulin, A., Grechka, V. & Tsvankin, I. 2001. Estimation of excavated rock slopes. PhD Thesis, University of London.
fracture parameters of monoclinic media from reflection Barton, N. 1972a. A model study of rock-joint deformation.
seismic data. 9th IWSA (Int. Workshop on Seismic Aniso- Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 9(5):
tropy), Abstracts, Geophysics, 66(4): 1294–1312. 579–602.
Bakulin, A., Grechka, V. & Tsvankin, I. 2002. Seismic inver- Barton, N. 1972b. A model study of air transport from under-
sion for the parameters of two orthogonal fracture sets in a ground openings situated below groundwater level. Proc. of
VTI background medium. Geophysics, 67: 292–299. ISRM Symp. on Percolation through Fissured Rock, Stuttgart,
Bamford, D. & Nunn, K.R. 1979. In-situ seismic measure- Theme 3A, 20 pp.
ments of crack anisotropy in the carboniferous limestone Barton, N. 1973a. Review of a new shear strength criterion
of Northwest England. Geophys. Prospect., 27: 322–338. for rock joints, Eng. Geol., 7: 287–332.
Bandis, S. 1980. Experimental studies of scale effects on shear Barton, N. 1973b. A review of the shear strength of filled dis-
strength, and deformation of rock joints. PhD Thesis, continuities in rock. Fjellspregningsteknikk, Bergmekanikk,
University of Leeds, England. 385 pp., and appendices. Oslo. 19.1–19.38 Trondheim: Tapir Press. Also NGI Publ.
Bandis, S., Lumsden, A. & Barton, N. 1981. Experimental 105, 1974.
studies of scale effects on the shear behaviour of rock Barton, N., Lien, R. & Lunde, J. 1974. Engineering classifi-
joints. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 18: cation of rock masses for the design of tunnel support.
1–21. Rock Mechanics, 6(4): 189–236, Springer Verlag, Vienna.
Bandis, S., Lumsden, A.C. & Barton, N. 1983. Funda- Barton, N. 1976. The shear strength of rock and rock joints.
mentals of rock joint deformation. Int. J. Rock Mech. Min. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 13(9):
Sci. & Geomech. Abstr., 20(6 ): 249–268. 255–279. Also NGI Publ. 119, 1978.
Bandis, S.C., Makurat, A. & Vik, G. 1985. Predicted and Barton, N. & Choubey, V. 1977. The shear strength of rock
measured hydraulic conductivity of rock joints. Proc. of joints in theory and practice. Rock Mechanics, 10(1/2):
Int. Symp. on Fundamentals of Rock Joints, Björkliden, 1–54. Vienna: Springer Verlag, Also NGI Publ. 119,
Sweden, 269–280. Luleå: CENTEK publishers. 1978.
Baquer, S. & Mitchell, B.J. 1998. Regional variation of Lg Barton, N. & Hansteen, H. 1979. Very large span openings
coda Q in the Continental United States and its relation to at shallow depth: deformation magnitudes from jointed
crustal structure and evolution. Pure Appl. Geophys., 153: models and F.E. analysis. Proc. of 4th Rapid Excavation and
613–638. Tunnelling Conf., Atlanta Georgia. Maevis & Hustrulid
Barazangi, M. & Isacks, B. 1971. Lateral variations of seis- (Eds.). 2: 1131–1353. New York: American Institute of
mic wave attenuation in the upper mantle above the Mining, Metallurgical, and Petroleum Engineers, Inc.
inclined earthquake zone of the Tonga Island arc: deep Barton, N. 1980. Discussion of paper by Krahn, J. &
anomaly in the upper mantle. J. Geophys. Res., 76: Morgenstern, N.R. The ultimate frictional resistance of
8493–8516. rock discontinuities. Int. J. Rock Mech. Min. Sci. &
Barclay, A.H. & Toomey, D.R. 2003. Shear wave splitting Geomech. Abstr., 17: 65–78.
and crustal anisotropy at the Mid-Atlantic ridge, 35° N. Barton, N. 1981. Hydraulic fracturing to estimate minimum
J. Geophys. Res., 108(B8): 2378, EM2-1–EM2-10. stress and rockmass stability at a pumped hydro project.
Barker, J.S., Hartzell, S.H., Burdick, L.J. & Helmburger, D.V. Proc. of Workshop on Hydraulic Fracturing Stress Measure-
1991. J. Geophys. Res., 96(B6 ): 10,129–10,143. ments, Monterey: CA.
References 629

Barton, N. 1982. Modelling rock joint behaviour from in waste repository design. 1992 Proc. of ISRM Symp. on
situ block tests: implications for nuclear waste repository EUROCK, Chester, UK.
design. Office of Nuclear Waste Isolation, Columbus: OH, Barton, N., Makurat, A., Monsen, K., Vik, G. & Tunbridge, L.
96 pp. ONWI-308, September 1982. 1992b. Geotechnical predictions of the excavation disturbed
Barton, N. & Bandis, S. 1982. Effects of block size on the zone at Stripa. Proc. of Fourth Int. Symp. on OECD/NEA
shear behaviour of jointed rock. 23rd US Symp. on Rock Stripa Project, 14–16 October 1992, Stockholm, Sweden.
Mechanics, Berkeley: CA. Goodman & Heuzé (Eds.). 77–96.
Keynote lecture. New York: Society of Mining Engineers. Barton, N. 1993. Physical and discrete element models of
Barton, N. & Bakhtar, K. 1983. Instrumentation and analy- excavation and failure in jointed rock. Keynote lecture
sis of a deep shaft in quartzite. Proc. of 24th US Symp. on presented at ISRM Int. Symp. on Assessment and Prevention
Rock Mechanics, Texas A&M University. of Failure Phenomena in Rock Engineering, 5–7 April,
Barton, N., Bandis, S. & Bakhtar, K. 1985. Strength, deforma- Istanbul, Turkey.
tion and conductivity coupling of rock joints. Int. J. Rock Barton, N. & Grimstad, E. 1994. The Q-system following
Mech. Min. Sci. & Geomech. Abstr., 22(3): 121–140. twenty years of application in NMT support selection.
Barton, N. 1986. Deformation phenomena in jointed rock. 43rd Geomechanic Colloquy, Salzburg. Felsbau, 6/94:
8th Laurits Bjerrum Memorial Lecture, Oslo. Geotechnique, 428–436.
36(2): 147–167. Barton, N., By, T.L., Chryssanthakis, P., Tunbridge, L.,
Barton, N., Hårvik, L., Christiansson, M., Bandis, S., Kristiansen, J., Løset, F., Bhasin, R.K., Westerdahl, H. &
Makurat, A., Chryssanthakis, P. & Vik, G. 1986. Rock Vik, G. 1994. Predicted and measured performance of the
mechanics modelling of the Ekofisk reservoir subsidence. 62 m span Norwegian Olympic ice hockey cavern at
Rock Mechanics: Key to Energy Production, Proc. of 27th Gjøvik. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr.,
U.S. Symp. Rock Mech., Tuscaloosa, AL, Hartman, H.L. 31(6 ): 617–641.
(Ed.). Society of Mining Engineers. Barton, N. 1995. The influence of joint properties in model-
Barton, N. 1987. Predicting the behaviour of underground ling jointed rock masses. Keynote lecture. Proc. of 8th ISRM
openings in rock. Manuel Rocha Memorial Lecture, Lisbon, Congress, Tokyo. T. Fuji (Ed.). 3: 1023–1032. Rotterdam:
NGI Publication 172, 1988. Also Geotecnia 53, July 1988 Balkema.
(in Portugese). Barton, N. & Itoh, J. 1995. The Q-system and NMT sup-
Barton, N., Makurat, A., Christianson, M. & Bandis, S. 1987. port techniques. (In Japanese) Tunnels and Underground,
Modelling rock mass conductivity changes in disturbed Japan Tunnelling Association Publication, November 95.
zones. Rock Mechanics. Proc. of 28th US Symp. Rock Barton, N. 1996a. Estimating rock mass deformation modu-
Mech., Tucson, AZ, Farmer, I.W., Daemen, J.J.K., Desai, lus for Excavation Disturbed Zone studies. Int. Conf. on
C.S., Glass, C.E. & Neuman, S.P. (Eds.): 563–574. Deep Geological Disposal of Radioactive Waste, Winnepeg,
Rotterdam: Balkema. 1996, EDZ Workshop, 133–144. Canadian Nuclear Society.
Barton, N., Makurat, A., Hårvik, L., Vik, G., Bandis, S., Barton, N. 1996b. Rock mass characterization and seismic
Christianson, M. & Addis, A. 1988. The discontinuum measurements to assist in the design and execution of TBM
approach to compaction and subsidence modelling as projects. Proc. of 1996 Taiwan Rock Engineering Symp.,
applied to Ekofisk. BOSS ’88. Proc. of Int. Conf. on Keynote Lecture, 1–16.
Behaviour of Offshore Structures, Trondheim. 1: 129–141. Barton, N. & Warren, C. 1996. Rock mass classification of
Barton, N. & Bandis, S.C. 1990. Review of predictive cap- chalk marl in the UK Channel Tunnels. Channel Tunnel
abilities of JRC-JCS model in engineering practice. Rock Engineering Geology Symp., Brighton, September 1995.
Joints, Barton, N. & Stephansson, O. (eds.). Int. Conf., Barton, N. & de Quados, E.F. 1997. Joint aperture and
Loen, Norway. 603–610. Rotterdam: Balkema. roughness in the prediction of flow and groutability of
Barton, N. 1990a. Chapter 6, panel of experts for Xiaolangdi rock masses. Proc. of NY Rocks ’97. Linking Science to
Project. Report No. 1 submitted to Yellow River Water & Rock Engineering. Kim, K. (Ed.). 907–916. Int. J. Rock
Hydroelectric Power Development Corporation, YRCC, Mech. Min. Sci. & Geomech. Abstr., 34: 3–4.
MWR. 26–42. Barton, N. 1998. Quantitative description of rock masses for
Barton, N. 1990b. Scale effects or sampling bias? Closing lec- the design of NMT reinforcement. Keynote Lecture, Int.
ture. Proc. of First Int. Workshop on Scale Effects in Rock Conf. on Hydro Power Development in Himalayas, Shimla,
Masses, Loen, Norway. Da Cunah (Ed.). LNEC. India. Rotterdam: Balkema.
Barton, N. 1991. Geotechnical Design. World Tunnelling, Barton, N. 1999a. General report concerning some 20th
November 1991. 410–416. Century lessons and 21st Century challenges in applied rock
Barton, N., Løset, F., Smallwood, A., Vik, G., Rawlings, C., mechanics, safety and control of the environment. Proc. of
Chryssanthakis, P., Hansteen, H. & Ireland, T. 1992a. 9th ISRM Congress, Paris, 3: 1659–1679. Rotterdam:
Geotechnical core characterization for the UK radioactive Balkema.
630 References

Barton, N. 1999b. Rock mass characterization from seismic in Eurasia. Scientific Report No. 1 to AFGL, December
measurements. SAROCKS ’98, 2nd Brazilian Symp. on 18, ENSCO, Inc. 90 pp.
5th South American Rock Mechanics Congress, Santos, Beaudoin, B.C., Fuis, G.S., Mooney, W.D. & Nokleberg, W.J.
Brazil. Keynote Lecture. 1992a. Thin, low-velocity crust beneath the Southern
Barton, N. 2000. TBM Tunnelling in Jointed and Faulted Yukon-Tanana Terrane, East Central Alaska: results from
Rock, 173 pp. Rotterdam: Balkema. Trans-Alaska crustal transect refraction/wide-angle reflec-
Barton, N. 2002a. Some new Q-value correlations to assist in tion data. J. Geophys. Res., 97(B2): 1921–1942.
site characterization and tunnel design. Int. J. Rock Mech. Beaudoin, B.C., ten Brink, U.S. & Stern, T.A. 1992b.
Min. Sci. & Geomech. Abstr., 39/2: 185–216. Characteristics and processing of seismic data collected on
Barton, N. 2002b. Different scales of discontinuous behav- thick floating ice: results from the Ross Ice Shelf, Antarctica.
iour in petroleum engineering. Keynote lecture, Proc. of Geophysics, 57(10): 1359–1372.
3rd Brazilian Rock Mechanics Symp., Sao Paulo, ABMS. Berge, P.A., Fryer, G.J. & Wilkens, R.H. 1992. Velocity-
Barton, N. & de Quadros, E. 2003. Improved understanding porosity relationships in the upper oceanic crust: theoretical
of high pressure pre-grouting effects for tunnels in jointed considerations. J. Geophys. Res., 97(B11): 15,239–15,254.
rock. Proc. of 10th ISRM Congress, South Africa, 1: 85–91, Bernabini, B. & Borelli, G.B. 1974. Methods for determining
Johannesburg: SAIMM. the average dynamic elastic properties of a fractured rock
Barton, N. 2004a. The theory behind high pressure grouting. mass and the variations of these properties near excavations.
Tunnels and Tunnelling International, 28–30 September, Proc. of 3rd ISRM Congress, Denver. IIA: 393–397.
October. 33–35. Washington, DC, National Academy of Sciences.
Barton, N. 2004b. Failure around tunnels and boreholes and Beroza, G.C., Cole, A.T. & Ellsworth, W.L. 1995. Stability
other problems in rock mechanics. Letter to ISRM News of coda wave attenuation during the Loma Prieta, California,
J., 8(2): May 2004, 12–18. earthquake sequence. J. Geophys. Res., 100: 3977–3987.
Barton, N. 2004c. Risk and risk reduction in TBM rock tun- Bertacchi, P., Carabelli, E. & Sampaolo, A. 1966. A contribu-
nelling. Keynote lecture, ARMS 2004: 3rd Asian Rock tion to the use of geophysical methods for investigation of
Mechanics Symp., Kyoto, Japan. rock masses. Proc. of 1st ISRM Congress, Lisbon. I: 45–53.
Barton, N. 2005. Seismic Q, rock quality, rock joint compliance LNEC.
and anisotropy. Seismic Q–observations, mechanisms and Bertacchi, P. & Sampaolo, A. 1970. Investigations on the char-
interpretation. Workshop 2, 67th EAGE Conf. on European acteristics of rock masses by geophysical methods. Proc. of
Association of Geoscientists and Engineers, Madrid. Extended 2nd ISRM Congress, Belgrade. III: Theme 8–13. Belgrade:
Abstract: Q-16, Convenors: Hargreaves. Best and Lancaster. Privredni Pregled.
Barton, N. & Makurat, A. 2006. Hydro-thermo-mechanical Best, A., Sothcott, J. & McCann, C. 2005. Laboratory ultra-
over-closure of joints and rock masses and potential effects on sonic Q anisotropy measurements on reservoir rocks. Seismic
the long term performance of nuclear waste repositories. Proc. Q–observation, mechanisms, and interpretation. Workshop at
of Eurock 06, Liege, Belgium. London: Taylor and Francis. 67th EAGE Conf., Madrid. Extended Abstract: Q-18,
Barton, P.J. 1986. The relationship between seismic velocity Convenors: Hargreaves, Best and Lancaster.
and density in the continental crust – a useful constraint? Best, A.I., McCann, C. & Sothcott, J. 1994. The relation-
Geophys. J. Roy. Astr. Soc., 878: 195–208. ships between the velocities, attenuations and petrophysi-
Batchelor, A.S. & Pine, R.J. 1986. The results of in situ stress cal properties of reservoir sedimentary rocks. Geophys.
determinations by seven methods to depths of 2500 m in Prospect., 42: 151–178.
the Carnmenellis granite. Proc. of Int. Symp. on Rock Stress Best, A.I. 1997. The effect of pressure on ultrasonic velocity
and Rock Stress Measurement, Stockholm, 467–478. and attenuation in near-surface sedimentary rocks. Geophys.
Batzle, M., Hoffmann, R., Kumar, G., Duranti, L., Prospect., 45: 345–364.
Han, D.-H. 2005. Seismic frequency loss mechanisms-direct Best, A.I. & Sams, M.S. 1997. Compressional wave velocity
observations. Q–observation, mechanisms, and interpretation. and attenuation at ultrasonic and sonic frequencies in near-
Workshop at 67th EAGE Conf. on European Association of surface sedimentary rocks. Geophys. Prospect., 45: 327–344.
Geoscientists and Engineers, Madrid. Extended Abstract: Bieniawski, Z.T. 1978. Determining rock mass deformabil-
Q-12, Convenors: Hargreaves, Best and Lancaster. ity: experience from case histories. Int. J. Rock Mech. Min.
Bauer, C., Homand-Etienne, F., Ben Slimane, K., Sci. & Geomech. Abstr., 15: 237–247.
Hinzen, K.G. & Reamer, S.K. 1996. Damage zone char- Bieniawski, Z.T. 1989. Engineering rock mass classifications:
acterization in the near field in the Swedish ZEDEX tun- A complete manual for engineers and geologists in mining,
nel using in situ and laboratory measurements. Eurock ’96. civil and petroleum engineering, 251pp. J. Wiley.
Barla (Ed.). 2: 1345–1352. Rotterdam: Balkema. Biot, M.A. 1956a. Theory of propagation of elastic waves in
Baumgardt, D.R. 1985. Attenuation, blockage and scatter- a fluid saturated solid, 1 lower frequency range. J. Acoust.
ing of teleseismic Lg from underground nuclear explosions Soc. Am., 28: 168–178.
References 631

Biot, M.A. 1956b. Theory of propagation of elastic waves in Bruce, B. & Bowers, G. 2002. Pore pressure terminology. The
a fluid saturated solid, 2 higher frequency range. J. Acoust. Leading Edge. February 2002: 170–173.
Soc. Am., 28: 179–191. Brueckl, E. & Parotidis, M. 2001. Estimation of large-scale
Biot, M.A. 1962. Generalized theory of acoustic propagation in mechanical properties of a large landslide on the basis
porous dissipative media. J. Acoust. Soc. Am., 34: 1254–1264. of seismic results. Int. J. Rock Mech. Min. Sci. & Geomech.
Birch, F. 1961. The velocity of compressional waves in rocks to Abstr., 38: 877–883.
10 kilobars. Part 2. J. Geoph. Res., 66(7 ): 2199– 2224. Bruno, M.S. & Winterstein, D.F. 1994. Some influences of
Boitnott, G.N., Biegel, R.L., Scholtz, C.H., Yoshioka, N. & stratigraphy and structure on reservoir stress orientation.
Wang, W. 1992. Micromechanics of rock friction Geophysics, 59(6 ): 954–962.
2: quantitative modelling of initial friction with contact Brzostowski, M.A. & McMechan, G.A. 1992. 3-D Tomo-
theory. J. Geophys. Res., 97(B6), 8965–8978. graphic imaging of near-surface seismic velocity and attenu-
Bonapace, B. 1983. Tests and measurements for the pressure ation. Geophysics, 57: 396–403.
tunnels and shafts of the Sellrain-Silz hydroelectric power Budianski, B. & O’Connell, R.J. 1976. Elastic moduli of a
scheme with extremely high head. Proc. of 5th Int. Congr. cracked solid. Int. J. Solids Struct., 12: 81–97, Elsevier.
Rock Mech. (ISRM), Melbourne: 2: D287–D292. Bull-Gjertsen, S.Y. 1998. New knowledge about Ekofisk –
Rotterdam: Balkema. four dimensional seismic gives improved reservoir control
Bott, M.H.P. 1982. The Interior of the Earth, Its Structure [in Norwegian]. Teknisk Ukeblad, 145(11): 34–35.
Constitution and Evolution, 2nd Edn. London: Edward By, T.L. 1987. Geotomography for rock mass characterization
Arnold. and prediction of seepage problems for two main road tun-
Bowers, G.L. 2002. Detecting high overpressure. The Leading nels under the city of Oslo. Proc. of 6th Int. Congr. Rock
Edge, February 2002, Soc. Explor. Geol., 174–177. Mech. (ISRM), Montreal, Herget, G. & Vongpaisal (Eds.):
Bradford, J.H. & Sawyer, D.S. 2002. Depth characterization of X(1): 37–40. Rotterdam: Balkema.
shallow aquifers with seismic reflection, Part II – Prestack By, T.L. 1988. Crosshole seismics including geotomography
depth migration and field examples. Geophysics, 67(1): for investigation of foundations. Int. Symp. on Field
98–109. Measurements in Geomechanics, 2. Kobe 1987. 1: 335–346.
Bradley, J.J. & Fort, Jr., A.N. 1966. Internal friction in rocks. Rotterdam: Balkema.
In: Handbook of physical constants: Clark, Jr., S.P. (Ed.). By, T.L., Lund, C. & Korhonen, R. 1988. Karst cavity detection
GSA Pub. 175–193. by means of seismic tomography to help assess possible dam
Bradley, W.B. 1978. Failure of inclined boreholes. Trans. Am. site in limestone terrain. Rock Mechanics and Power Plants,
Soc. Mech. Eng., 101: 232–239. Romana (Ed.). 1: 25–31. Rotterdam: Balkema.
Brandt, H. 1955. A study of the speed of sound in porous gran- Byerlee, J.D. 1978. Friction of rocks. Pure Appl. Geophy.,
ular media. J. Appl. Mech. ASME, December 479–486. 116: 615–629.
Brennan, B.J. & Stacey, F.D. 1977. Frequency dependence of
elasticity of rock – Test of seismic velocity dispersion. Nature,
268: 220–222. C
Brie, A., Endo, T., Hoyle, D., Codazzi, D., Esmersoy, C.,
Hsu, K., Denoo, S., Mueller, M.C., Plona, T., Shenoy, R. & Cadoret, T. 1993. Effect de la saturation eau/gaz sur les pro-
Sinha, B. 1998. New Directions in Sonic Logging. prieties acoustics des roches. PhD thesis, University of
Oilfield Rev., 40–55. Spring 1998, Schlumberger. Paris.
Briggs, V., Rama Rao, V.N., Grandi, S., Burns, D. & Chi, S. Cai, M., Kaiser, P.K. & Martin, C.D. 1998. A tensile model
2004. Research Report, MIT Earth Resources Lab., 12 pp. for the interpretation of microseismic events near under-
Brocher, T.M. & ten Brink, U.S. 1987. Variations in ground openings. Pure Appl. Geophys., 153: 67–92.
oceanic layer 2 elastic velocities near Hawaii and their Capozza, F. 1979. Panel discussion: dynamic testing. Int.
correlation to lithospheric flexure. J. Geophys. Res., 92(B3): Symp. on The Geotechnics of Structurally Complex
2647–2661. Formations. Capri 1977. II: 275–279. Associazione
Brown, A. 2001. Understanding seismic attributes. Geophysics, Geotecnica Italiana.
66: 47–48. Carcione, J.M. 1992. Anisotropic Q and velocity dispersion
Brown, R.L. & Siefert, D. 1997. Velocity dispersion: a tool for of finely layered media. Geophys. Prospect., 40: 761–783.
characterizing reservoir rocks. Geophysics, 62: 447–486. Carcione, J.M. 2000. A model for seismic velocity and atten-
Brown, R.L., Penuela, G., Hughes, R.G., Wiggens, M.L., uation in petroleum source rocks. Geophysics, 65(4 ):
Civan, F. & Martinez Torres, L. 2002. Relating saturation 1080–1092.
and permeability to elastic properties of fractured rocks. Carcione, J.M. & Gangi, A.F. 2000a. Gas generation and
Society of Exploration Geophysicists, Expanded Abstracts: overpressure: effects on seismic attributes. Geophysics, 65:
113–116. 1769–1779.
632 References

Carcione, J.M. & Gangi, A.F. 2000b. Non-equilibrium com- Chapman, M. & Liu, E. 2004. Theories for fluid substitu-
paction and abnormal pore-fluid pressures: effects on rock tion in fractured carbonate reservoirs. EAGE 66th Conf.
properties. Geophys. Prospect., 48: 521–537. on Exhibition, Paris.
Carcione, J.M. & Cavallini, F. 2002. Poisson’s ratio at high Chen, H., Brown, R.L. & Castagna, J.P. 2005. AVO for
pore pressure. Geophys. Prospect., 50: 97–106. one- and two-fracture set models. Geophysics, 70(2):
Carlson, R.L. 1998. Seismic velocities in the uppermost C1–C5.
oceanic crust: age dependence and the fate of layer 2A. Chen W.-H. 1982. Classification of soft wall rock of the Xia-
J. Geophys. Res., 103: 7069–7077. Keng tunnel on the Anhui-Jiangxi railway. Proc. of 4th Int.
Carlson, S.R. & Young, R.P. 1993. Acoustic emission and Congr. Int. Ass. Eng. Geol. (IAEG), New Delhi: V:
ultrasonic velocity study of excavation-induced microcrack V.241–V.246. Rotterdam: Balkema.
damage at the Underground Research Laboratory. Int. J. Chen, W.-Y., Lovell, C.W., Haley, G.M. & Pyrak-Nolte, L.J.
Rock Mech. Min. Sci. & Geomech. Abstr., 30(7 ): 901–907. 1993. Variation of shear-wave amplitude during frictional
(34th US Symp. on Rock Mechanics, Madison, WI.) sliding. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr.,
Carpenter, P.J. & Sanford, A.R. 1985. Apparent Q for upper 30(7 ): 779–784. (34th US Symp. on Rock Mechanics,
crustal rocks of the Central Rio Grande Rift. J. Geophys. Madison: WI.)
Res., 90: 8661–8674. Chen, X. & Long, L.T. 2000. Hypocenter migration as an
Carrère, A. 1999. Application of the SCARABEE method to explanation for temporal changes in coda Q. J. Geophys.
re-evaluation of concrete dams safety. Geodia report. France: Res., 105: 16,151–16,160.
Coyne et Bellier. Chi, S., Wu, J. & Torres-Verdin, C. 2002. Invasion correc-
Carter, A.J. & Kendall, J.-M. 2005. Differential attenuation of tion of Acoustic Logs in a Gas Reservoir. 43rd Annual
‘split’ shear waves – a future tool for reservoir charac- Symp. on SPWLA, Osio, Japan. Extended Abstracts.
terization? Seismic Q–observation, mechanisms, and interpre- Chi, S., Torres-Verdin, C., Wu, J. & Alpak, O.F. 2004.
tation. Workshop at 67th EAGE Conf., Madrid. Extended Assessment of mud-filtrate invasion effects on borehole
Abstract: Q-11, Convenors: Hargreaves, Best and Lancaster. acoustic logs and radial profiling of formation elastic
Castagna, J.P., Batzle, M.L. & Kan, T.K. 1993. Rock physics – parameters. SPE, Paper 90159, 13 pp.
the link between rock properties and AVO response. In: Christensen, N.I. & Salisbury, M.H. 1972. Sea floor spread-
Offset-Dependent Reflectivity – Theory and Practice of AVO ing, progressive alteration of layer 2 basalts, and associated
Analysis, Castagna, J.P. & Bakus, M. (Eds.). Investigations changes in seismic velocities. Earth Planet, Sci. Lett., 15:
in Geophysics, No. 8: 135–171. Tulsa, OK: Society of 367–375.
Exploration Geophysicists. Christensen, N.I. 1984. Pore pressure and oceanic crustal
Castagna, J.P. 2001. Recent advances in seismic lithologic seismic structure. Geophys. J. Roy. Astr. Soc., 79: 411–423.
analysis. Geophysics, 66: 42–46. Christensen, N.I. & Mooney, W.D. 1995. Seismic velocity
Cecil, O.S. 1971. Correlation of seismic refraction velocities structure and composition of the continental crust: a global
and rock support requirements in Swedish tunnels. Statens view. J. Geophys. Res., 100(B7): 9761–9788.
Geotekniska Institut. Publ. 40. 58 pp. Christeson, G.L., Purdy, G.M. & Fryer, G.J. 1991. Structure
Chan, T.C.F. 1993. Study of a proposed underground refuse of young upper crust at the East Pacific Rise near
transfer station in Hong Kong. 34th US Symp. on Rock 9°30 N. Geophys. Res. Lett., 19: 1045–1048.
Mechanics, Madison: WI. Haimson (Ed.). II: 773–776. Christeson, G.L., Wilcock, W.S.D. & Purdy, G.M. 1994. The
University of Wisconsin. shallow attenuation structure of the fast-spreading East
Chang, H. & Lee, T.S. 2001. Current use of geophysical site Pacific Rise near 9°30N. Geophys Res. Lett., 21: 321–324.
characterization techniques for civil engineering in Korea. Christie, P., Ireson, D., Rutherford, J., Smith, N., Dodds, K.,
Proc. of 4th Int. Workshop on Application of Geophysics to Johnston, L. & Schaffner, J. 1995. Borehole seismic data
Rock Engineering, Beijing, China. Int. Soc. Rock Mech. sharpen the reservoir image. Oilfield Rev., Schlumberger.
News J., 7(1): 45–52. Chryssanthakis, P., Monsen, K. & Barton, N. 1991. Vali-
Chapman, M., Zatsepin, S.V. & Crampin, S. 2001. On fre- dation of UDEC-BB against the CSM block test and large
quency scaling in anisotropic poro-elasticity. 9th IWSA scale application to glacier loading of jointed rock masses.
(Int. Workshop on Seismic Anisotropy), Abstracts, Geophysics, Proc. of 7th ISRM Cong., Aachen. 693–698. Rotterdam:
66(4): 1294–1312. Balkema.
Chapman, M. 2002. Frequency dependent anisotropy due to Chuhan, F.A. & Bjørlykke, K. 2002. Evidence of grain
meso-scale fractures with equant porosity. EAGE 64th crushing during mechanical compaction of reservoir sand-
Conf. and Exhibition, Florence, Italy. stones – A comparison with experimental results.
Chapman, M. 2003. Frequency-dependent anisotropy due American Association of Petroleum Geologists. (SEPM:
to meso-scale fractures in the presence of equant porosity. Effects of Deep Burial on Reservoir and Seal Quality),
Geophys. Prospect., 51: 369–379. Annual Meeting, Houston, TX.
References 633

Clouser, R.H. & Langston, C.A. 1991. Qp Qs relations in a Crampin, S. & Booth, D.C. 1985. Shear-wave polarizations
sedimentary basin using converted phases. Bull. Seism. Soc. near the North Anatolian fault – II interpretation in terms of
Am., 81: 733–750. crack-induced anisotropy. Geophys. J. Roy. Astr. Soc., 83:
Cole, K.S. & Cole, R.H. 1941. Dispersion and absorption in 75–92.
dielectrics. J. Chem. Phys., 9: 341–351. Crampin, S. 1989. Suggestions for a consistent terminology
Collier, J.S. & Singh, S.C. 1998. Poisson’s ratio structure for seismic anisotropy. Geophys. Prospect., 37: 753–770.
of young oceanic crust. J. Geophys. Res., 103(B9): Crampin, S. & Booth, D.C. 1989. Shear-wave splitting
20,981–20,996. showing hydraulic dilation of pre-existing joints in gran-
Comité National Français 1964. La deformabilite des massifs ite. Sci. Drill., 1: 21–26.
rocheux. Analyse et comparaison des resultats. Int. Congress Crampin, S. & Lovell, J.H. 1991. A decade of shear-wave split-
on Large Dams, 8. Edinburgh 1964. 1 (Q28): 287–299. ting in the Earth’s crust: what does it mean? What use can
Paris: ICOLD. we make of it? And what should we do next? Geophys. J.
Cooke, N.G.W. 1992. Natural joints in rock: mechanical, Int., 107: 387–407.
hydraulic and seismic behaviour and properties under nor- Crampin, S. 1993a. Arguments for EDA. Can. J. Explor.
mal stress. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., Geophys., 29(1): 18–30.
29 : 198–223. Crampin, S. 1993b. A review of the effects of crack geometry
Cosma, C. 1983. Determination of rock mass quality by the on wave propagation through aligned cracks. Can. J.
crosshole seismic method. Bull. Int. Ass. Eng. Geol., Explor. Geophys., 29(1): 3–17.
26–27 : 219–225. Crampin, S. 2000. Shear-wave splitting in a critical self-
Cosma, C. 1995. Characterization of subsurface structures organised crust: the new geophysics. 70th. Annual Int.
by remote sensing. Keynote lecture. Proc. of 8th ISRM Meeting Soc. Explor. Geophys., Calgary.
Congress, Tokyo. Fuji (Ed.). 3: 1013–1021. Rotterdam: Crampin, S., Volti, T., Chastin, S., Gudmundsson, A. &
Balkema. Stefánsson, R. 2002. Indication of high pore-fluid
Cosma, C. & Enescu, N. 1996a. Detailed seismic tomogra- pressures in a seismically-active fault zone. Geophys. J. Int.,
phy in radial boreholes – D&B. Technical Note. 151: F1–F5.
Stockholm: SKB. Crampin, S. 2003. The new geophysics: shear-wave splitting
Cosma, C. & Enescu, N. 1996b. Detailed seismic tomogra- provides a window into the crack-critical rock mass. The
phy in radial boreholes – TBM. Technical Note. Leading Edge, June 2003.
Stockholm: SKB. Crampin, S. & Peacock, S. 2005. A review of shear-wave
Cosma, C., Olsson, O., Keskinen, J. & Heikkinen, P. 2001. splitting in the compliant crack-critical anisotropic earth.
Seismic characterization of fracturing at the Äspö Hard Wave Motion, 41: 59–77.
Rock Laboratory, from the kilometre scale to the meter Cumerlato, C.L., Stachura, V.J. & Tweeton, D.R. 1988.
scale. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 38: Application of refraction tomography to map the extent of
859–863. blast-induced fracturing. 29th US Symp. on Rock
Coyne et Bellier & Geodia, 1995. SCARABEE – Mechanics, Minneapolis: MN. Cundall et al. (Eds.).
Investigation method for the rapid assessment of rock 691–698. Rotterdam: Balkema.
mass quality. April 1995. France: Coyne et Bellier. Cundall, P.A. 1971. The measurement and analysis of accelera-
Coyne et Bellier, 1998. Application de la methode tions in rock slopes. PhD Thesis, University of London,
SCARABEE. Exploitation de la base èxpérimentale. France: Imperial College, 182 pp.
Coyne et Bellier.
Crampin, S. 1978. Seismic waves propagating through a
cracked solid: polarization as a possible dilatancy diagnos- D
tic. Geophys. J. Roy. Astr. Soc., 53: 467–496.
Crampin, S., McGonigle, R. & Bamford, D. 1980. Daly, R.A. 1940. Strength and Structure of the Earth, 434 pp.,
Estimating crack parameters from observations of P-wave Englewood Cliffs, NJ: Prentice-Hall.
velocity anisotropy. Geophysics, 45(3): 345–360. Dangerfield, J.A. & Brown, D.A. 1985. The Ekofisk Field:
Crampin, S. 1981. A review of wave motion in an isotropic North Sea Oil and Gas Reservoirs, Kleepe, J. et al. (Eds.):
and cracked elastic-medium. Wave Motion, 3: 343–391. 3–22, London: Graham & Trotman.
Crampin, S. 1984. Effective anisotropic elastic constants for Darracott, B.W. & Orr, C.M. 1976. Geophysics and rock
wave propagation through cracked solids. Geophys. J. Roy. engineering. Symp. on Exploration for Rock Engineering.
Astr. Soc., 76: 135–145. Johannesburg 1976. 1: 159–164. Cape Town/Rotterdam:
Crampin, S., Evans, R. & Atkinson, B.K. 1984. Earthquake Balkema.
prediction: a new physical basis. Geophys. J. Roy. Astr. Soc., Dasgupta, R. & Clarke, R.A. 1998. Estimation of Q from sur-
76: 147–156. face seismic reflection data. Geophysics, 63(6 ): 2120–2128.
634 References

Dasios, A., McCann, C., Austin, T.R., McCann, D.M. & Dvorkin, J. & Nur, A. 1993. Dynamic poroelasticity: a uni-
Fenning, P. 1999. Seismic imaging of the shallow subsur- fied model with the squirt and the biot mechanism. Geo-
face: shear wave case histories. Geophys. Prospect., 47: physics, 58: 524–533.
565–591. Dvorkin, J., Nolen-Hoeksema, R. & Nur, A. 1994. The squirt
Dasios, A., Astin, T. & McCann, C. 2001. Compressional- flow mechanism: macroscopic description. Geophysics, 59:
wave Q estimation from full waveform sonic data. Geophys. 428–438.
Prospect., 49: 353–373. Dvorkin, J., Mavko, G. & Nur, A. 1995. Squirt flow in fully
Davies, T.L. 2001. Multicomponent seismiology – the next saturated rocks. Geophysics, 60: 97–107.
wave. Geophysics, 66: 49. Dvorkin, J. & Nur, A. 1996. Elasticity of high porosity sand-
De Souza, J.L. & Mitchell, B.J. 1998. Lg Coda Q variations stones: theory for two North Sea data sets. Geophysics, 61:
across South America and their relation to crustal evolu- 1363–1370.
tion. Pure Appl. Geophys., 153: 587–612. Dvorkin, J. & Nur, A. 1998. Short Note: time-average equa-
De, G.S., Winterstein, D.F. & Meadows, M.A. 1994. tion revisited. Geophysics, 63: 460–464.
Comparison of P- and S-wave velocities and Q’s from VSP Dyer, B.C., Jones, R.H., Cowles, J.F. & Barkved, O. 1999.
and sonic log data. Geophysics, 59: 1512–1529. Microseismic monitoring of a North Sea reservoir. World
Deere, D.U., Hendron, A.J., Patton, F.D. & Cording, E.J. Oil, 220(3): 73–78.
1967. Design of surface and near-surface construction. In: Dypvik, H., Gudlaugsson, S.T., Tsikalas, F., Attrep, Jr., M.,
Rock. Failure and Breakage of Rock, Fairhurst, C. (Ed.): Ferrell, Jr., R.E., Krinsley, D.H., Mørk, A., Faleide, J.I. &
237–302. New York: Society of Mining Engineers of AIME. Nagy, J. 1996. Mjølnir structure: an impact crater in the
Den Boer, L. & Matthews, L. 1988. Seismic characterization Barents Sea. Geology, 24(9): 779–782.
of thermal flood behaviour. Unitar/UNDP Fourth Int. Conf.
on Heavy Crude and Tar Sands.
Denekamp, Saul. Personal communication. E
Detrick, R.S., White, R.S. & Purdy, G.M. 1993. Crustal
structure of North Atlantic fracture zones. Rev. Geophys., Ebisu, S., Aydan, Ö., Komura, S. & Kawamoto, T. 1992.
31(4): 439–458. Comparative study on various rock mass characterization
Dhawan, A.K., Kate, J.M. & Joshi, A.B. 1983. Geophysical methods for surface structures. ISRM Symp. on Eurock ’92.
exploration for a project in the Himalayas. Proc. of 5th ISRM Chester, UK: 203–208. London: Thomas Telford.
Congress, Melbourne. 1: A9–A13. Rotterdam: Balkema. Ecevitoglu, B. & Bingol, H. 1999. Q imaging from first
Di Salvo, C.A. 1982. Geomechanical classification of rock breaks in shallow seismics. 2nd Balkan Geophysics Cong. on
mass at Segunda Angostura dam, Limay river, Argentine Exhib., Extended Abstract: O12-3, 110–111.
Republic. Int. Congress on Large Dams, 14. Rio de Janeiro. Eide, A.L., Alsos, T., Hegstad, B.K., Najjar, N.F., Astratti, D.,
2(Q53): 519–530. Paris: ICOLD. Psaila, D. & Doyen, P. 2002. Quantitative time-lapse seis-
Domenico, S.N. 1984. Rock lithology and porosity determi- mic analysis of the Gullfaks field. Geophysics in a shared
nation from shear and compressional wave velocity. earth model. Symp., Norwegian Petroleum Society NPF,
Geophysics, 49: 1188–1195. Kristiansand, Extended Abstract: 61–64.
Domnesteanu, P., McCann, C. & Sothcott, J. 2002. Velocity Emsley, S.J., Olsson, O., Stanfors, R., Stenberg, L.,
anisotropy and attenuation of shale in under- and over- Cosma, C. & Tunbridge, L. 1996. Integrated characteriza-
pressured conditions. Geophys. Prospect., 50 : 487–503. tion of a rock volume at the Äspö HRL utilised for an
Dorman, L.M. & Jacobson, R.S. 1981. Linear inversion of EDZ experiment. Eurock ’96. Barla (Ed.). 2: 1329–1336.
body wave data, Part 1: Velocity structure from travel times Rotterdam: Balkema.
and ranges. Geophysics, 46(2): 138–151. Endo, T., Ito, H., Badri, M. & El Sheikh, M. 1997. Fracture
Duellmann, H. & Heitfeld, K.H. 1978. Influence of grain and permeability evaluation in a fault zone from sonic
fabric anisotropy on the elastic properties of rocks. Proc. of waveform data. Trans. SPWLA 38th Ann. Logging Symp.,
3rd IAEG Congress, Madrid. II(1): 150–162. Madrid: Houston, Paper R.
Imprime ADOSA. Engelder, T. & Plumb, R. 1984. Changes in in situ ultrasonic
Dürrast, H., Rasolofosaon, P.N.J. & Siegesmund, S. 2002. properties of rock on strain relaxation. Int. J. Rock Mech.
P-wave velocity and permeability distribution of sand- Min. Sci. & Geomech. Abstr., 21(2): 75–82.
stones from a fractured tight gas reservoir. Geophysics, 67(1): Esaki, T., Ikusada, K., Aikawa, A. & Kimura, T. 1995. Surface
241–253. roughness and hydraulic properties of sheared rock. Int.
Dusseault, M.B., Bruno, M.S. & Barrera, J. 2001. Casing Conf. on Fractured and Jointed Rock Masses. Myer, Tsang,
shear: causes, cases, cures. SPE Paper 72060, June 2001, Cook & Goodman (Eds.): 93–98. Lake Tahoe, CA.
SPE Drilling and Completion, 98–107. Tulsa, OK: Society Evangelista, A. & Pellegrino, A. 1990. Caratteristiche geot-
of Petroleum Engineers. echiche di alcune rocce tenere Italiane. Conferenze di
References 635

Meccanica e Ingegneria delle Rocce. Politecnico de Torino conditions: observations and implications. Int. J. Rock Mech.
1990. Terzo Ciclo. 2-1-2-32. Padova: SGE Editoriali. Min. Sci. & Geomech. Abstr., 33(3): 279–290.
Evans, R.L. 1994. Constraints on the large-scale porosity and Friedel, M.J., Scott, D.F., Jackson, M.J., Williams, T.J. &
permeability structure of young oceanic crust from velocity Killen, S.M. 1996b. 3-D tomographic imaging of
and resistivity data. Geophys. J. Int., 119: 869– 879. anomalous stress conditions in a deep US gold mine.
Ewing, J. 1963. Elementary theory of seismic refraction and J. Appl. Geophys., 36: 1–17.
reflection measurements. In: The Sea, Vol. 3, Hill, M.N. Frisillo, A.L. & Stewart, J. 1980. Effect of partial gas/brine
(Ed.): 3–19. New York: Interscience. saturation on ultrasonic absorption in sandstone. J.
Ewing, M. 1965. The sediments of the Argentine basin. Q. J. Geophys. Res., 85: 5209–5211.
Roy. Astr. Soc., 6: 10–27.
Ewing, J.I. & Purdey, G.M. 1982. Upper crustal velocity struc-
ture in the ROSE area of the East Pacific Rise. J. Geophys. G
Res., 87(B10): 8397–8402.
Gaiser, J. & van Dok, R. 2003. Converted shear-wave
anisotropy: new technology for fractured-reservoir man-
F agement. AAPG Int. Conf., Barcelona, Spain.
Gale, J.E. 1987. Comparison of coupled fracture deform-
Faust, L.Y. 1951. Seismic velocity as a function of depth and ation and fluid flow models with direct measurements
geologic time, Geophysics 16(2): 192–206. of fracture pore structure and stress-flow properties.
Fessenden, R.A. 1917. Method and apparatus for locating Rock Mechanics, Proc. of 28th US Symp. Rock Mech.,
ore bodies. US patent 1.240,328. Tucson, AZ, Farmer, I.W., Daemen, J.J.K., Desai, C.S.,
Feustel, A.J., Trifu, C.-Z. & Urbancic, T.i. 1996. Rock mass Glass, C.E. & Neuman, S.P. (Eds.): 1213–1222.
characterization using intrinsic and scattering attenuation Rotterdam: Balkema.
estimates at frequencies from 400 to 1600 Hz. Pure Appl. Gangi, A.F. 1991. The effect of pore fluids and pressures on
Geophys., 147(2): 289–304. the seismic velocities in cracked and/or porous rocks. SEG
Fjær, E., Holt, R.M., Horsrad, P., Raaen, A.M. & Risnes, R. Research Workshop on Lithology Tech, Abstracts, 35–38.
1992. Petroleum Related Rock Mechanics, Elsevier. Gardner, G.H.F., Gardner, L.W. & Gregory, A.R. 1974.
Fjær, E. & Holt, R.M. 1994. Rock acoustics and rock Formation velocity and density – the diagnostic basics for
mechanics: their link in petroleum engineering. The stratigraphic traps. Geophysics, 39(6 ): 770–780.
Leading Edge, April 1994, 255–258. Gardener, R. 1992. Seismic refraction as a tool in the evalu-
Fjær, E. 1997. Laboratory tests on artificial rocks with con- ation of rock quality for dredging and engineering
trolled crack parameters. Int. J. Rock Mech. Min. Sci. & purposes: case studies. ISRM Symp. on Eurock ’92. Chester,
Geomech. Abstr., 34: 3–4. UK: 153–158. London: Thomas Telford.
Fjær, E. & Suárez-Rivera, R. 1998. Fracture size determined Geertsma, J. 1961. Velocity-log interpretation: the effect of
from amplitude data. EAGE 60th Conf. on Technical rock bulk compressibility. Soc. Petrol. Eng. J., I: 235–248.
Exhibition, Leipzig, Germany. Gentier, S. 1986. Morphologie et comportement hydro-
Fjær, E. & Holt, R.M. 1999. Stress and stress release effects mechanique d’une fracture naturelle dans un granite sous
on acoustic velocities from cores, logs and seismics. SPWLA contrainte normale. PhD Thesis, Universite d’Orleans, France.
40th Ann. Logging Symp., Paper WW. Geo, 2001. Ekofisk – old field towards new heights. (in
Fourmaintraux, D. 1995. Quantification des discontinuités Norwegian) Geo- Bergverksnytt, February 2001, 12–16.
des roches et des massifs rocheux. Rock Mech., 7: 83–100. Trondheim.
Fratta, D. & Santamarina, J.C. 2002. Shear wave propagation Gibbs, J.F. & Roth, E.F. 1989. Seismic velocities and atten-
in jointed rock: state of stress. Géotechnique, 52(7): 495–505. uation from borehole measurements near the Parkfield
Frazer, L.N., Sun, X. & Wilkens, R.H. 1997. Changes in prediction zone. Central California. Earthq. Spectr., 5:
attenuation with depth in an ocean carbonate section: 513–537.
Ocean Drilling Program sites 806 and 807, Ontong Java Gibson, R.L. & Toksöz, M.N. 1990. Permeability estimation
Plateau. J. Geophys. Res., 102(B2): 2983–2997. from velocity anisotropy in fractured rock. J. Geophys. Res.,
Friedel, M.J., Scott, D.F., Jackson, M.J., Williams, T.J. & Killen, 95(B10): 15,643–15,655.
S.M. 1995. 3D tomographic imaging of mechanical condi- Goldberg, D. & Yin, C.-S. 1994. Attenuation of P waves in
tions in a deep US gold mine. Mechanics of jointed and oceanic crust: multiple scattering from observed hetero-
faulted rock. Int. Conf. on Mechanics of Jointed and Faulted geneities. Geophys. Res. Lett., 21: 2311–2314.
Rock, 2., Vienna: 689–695. Rotterdam: Balkema. González, M. & Munguía, L. 2003. Seismic anisotropy
Friedel, M.J., Jackson, M.J., Williams, E.M., Olson, M.S. & observations in the Mericali Valley, Baja California, México.
Westman, E. 1996a. Tomographic imaging of coal pillar Pure Appl. Geophys., 160: 2257–2278.
636 References

Goodman, R.E., Taylor, R.L. & Brekke, T.L. 1968. A model Industry, Proc. of 1st North Am. Rock Mech. Symp., Austin,
for the mechanics of jointed rock. J. Soil Mech. Found. TX, Nelson, P. & Laubach, S.E. (Eds.): 165–172.
Div., ASCE, 94: 637–659. Rotterdam: Balkema.
Goodman, R.E. 1970. The deformability of joints. Am. Soc. Gutierrez, M. & Makurat, A. 1997. Coupled HTM modell-
Test Materials, Special Test Publication. 477: 174–196. ing of cold water injection in fractured hydrocarbon reser-
Goodman, R.E. 1976. Methods of Geological Engineering in voirs. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 34:
Discontinuous Rocks. 172 pp. New York: West Publishing. 3–4, Paper 113.
Gordon, R.B. & Davis, L.A. 1968. Velocity and attenuation Gutierrez, M.A., Dvorkin, J. & Nur, A. 2002. Stratigraphy-
of seismic waves in imperfectly elastic rock. J. Geophys. guided rock physics. The Leading Edge, January 2002,
Res., 73: 3917–3935. 98–103.
Grainger, P., McCann, D.M. & Gallois, R.W. 1973. The
application of the seismic refraction technique to the study
of fracturing of the Middle Chalk at Mundford, Norfolk. H
Geotechnique, 23: 219–232.
Granger, P.-Y., Rollet, A. & Bonnot, J.-M. 2000. Preliminary Hackert, C.L., Parra, H.O., Brown, R.L. & Collier, H.A.
evaluation of azimuthal anisotropy over the Valhall field 2001. Characterization of dispersion, attenuation, and
using C-wave data. Anisotropy 2000: fractures, converted anisotropy at the Buena Vista Hills Field. Geophysics, 66:
waves, and case studies. Society of Exploration Geophysicists, 90–96.
772–775. Hackert, C.L. & Parra, J.O. 2003. Estimating scattering from
Green, A.S.P., Baria, R. & Jones, R. 1989. VSP and cross- vugs or karsts. Geophysics, 68(4 ): 1182–1188.
hole seismic surveys used to determine reservoir character- Haimson, B.C. 2006. Borehole breakouts in Berea Sandstone
istics of a hot dry rock geothermal system. Int. J. Rock reveal a new fracture mechanism. Pure Appl. Geophys., 160:
Mech. Min. Sci. & Geomech. Abstr., 26(3/4): 271–280. 813–831.
Gregory, A.R. 1976. Fluid saturation effects on dynamic elas- Halderman, T.P. & Davis, P.M. 1991. Qp beneath the Rio
tic properties of sedimentary rocks. Geophysics, 41: Grande and East African rift zones. J. Geophys. Res., 96:
895–923. 10,113–10,128.
Grevemeyer, I., Kaul, N. & Villinger, H. 1999. Hydrothermal Hall, S.A. & Kendall, J.-M. 2000. Constraints on the inter-
activity and the evolution of the seismic properties of upper pretation of P-wave AVAO for fracture characterization.
oceanic crust. J. Geophys Res., 104(B3): 5069–5079. Proc. of 9th Int. Workshop on Seismic Anisotropy (IWSA),
Griffiths, D.H. & King, R.F. 1987. Geophysical exploration. Houston, TX: Extended Abstract. Society of Exploration
Ground Engineers Reference Book. Bell, F.G. (Ed.). London: Geophysicists.
Butterworths, 27/1–27/21. Hall, S.A. & Kendall, J.-M. 2003. Fracture characterization
Grimm, R.E., Lynn, H.B., Bates, C.R., Phillips, D.R., at Valhall: application of P-wave amplitude variation with
Simon, K.M. & Beckham, W.E. 1999. Detection and analy- offset and azimuth (AVOA) analysis to a 3D ocean-bottom
sis of naturally fractured gas reservoirs: multiazimuth seismic data set. Geophysics, 68(4 ): 1150–1160.
surveys in the Wind River basin, Wyoming. Geophysics, Hamilton, E.L. 1971. Elastic properties of marine sedi-
64(4): 1277–1292. ments. J. Acoust. Soc. Am., 28: 266–284.
Grujic, N. 1974. Ultrasonic testing of foundation rock. Proc. Hamilton, E.L. 1972. Compressional-wave attenuation in
of 3rd ISRM Congress, Denver. IIA: 404–409. Washington, marine sediments. Geophysics, 37: 620–646.
DC: National Academy of Sciences. Hamilton, E.L. 1979. Vp/Vs and Poissons’ ratios in marine sed-
Guilbot, J., Smith, B. & Pirera, F. 2002. Using 4D seismic to iments and rocks. J. Acoust. Soc. Am., 66(4 ): 1093– 1101.
understand the complex relationship between water injec- Han, D. 1986. Effects of porosity and clay content on acoustic
tion and reservoir compaction at ekofisk field. Geophysics properties of sandstones and unconsolidated sediments.
in a Shared Earth Model. Symp., Norwegian Petroleum PhD Thesis, Stanford University, USA.
Society NPF, Kristiansand, Extended Abstract: 45–52. Han, D., Nur, A. & Morgan, D. 1986. Effects of porosity
Gurevich, B. 2002. Effect of the pore fluid on the elastic prop- and clay content on wave velocities in sandstones.
erties of a porous rock with parallel fractures. 10th Int. Geophysics, 51: 2093–2107.
Workshop on Seismic Anisotropy. Evangelische Akademie, Hardin, E.L., Barton, N., Lingle, R., Board, M.P. &
Tutzing, Germany, Prgm. and Abstracts, 72 pp. Voegele, M.D. 1981. A heated flatjack test series to measure
Gurevich, B. 2003. Elastic properties of saturated porous rocks the thermomechanical and transport properties of in situ
with aligned fractures. J. Appl. Geophys., 54: 203–218. rock masses. Office of Nuclear Waste Isolation, Columbus:
Gutierrez, M. & Barton, N. 1994. Numerical Modelling of OH. ONWI-260, 193 pp.
the Hydro-Mechanical Behavior of Single Fractures. Rock Harding, A.J., Orcutt, J.A., Kappus, M.E., Vera, E.E.,
Mechanics Models and Measurements: Challenges from Mutter, J.C., Buhl, P., Detrick, R.S. & Brocher, T.M.
References 637

1989. Structure of young oceanic crust at 13°N on the Heimli, P. 1972. Water and air leakage through joints in rock
East Pacific Rise from expanding spread profiles. J. Geophys. specimens. (in Norwegian). Fjellsprengningsteknikk,
Res., 94: 12,163–12,196 Bergmekanikk. Annual Rock Mechanics Meeting, Oslo,
Harjes, H.-P., Brann, K., Dürbaum, H.-J., Gebrande, H., Norway, 137–142, Ingeniørforlaget, Oslo.
Hirschmann, G., Janik, M., Klöckner, M., Lüschen, E., Helbig, K. & Szaraniec, E. 2000. The historical roots of
Rabbel, W., Simon, M., Thomas, R., Tormann, J. & anisotropy. Proc. of 9th Int. Workshop on Seismic Anisotropy
Wenzel. 1997. Origin and nature of crustal reflections: (IWSA), Houston, TX: Extended Abstract. Society of
results from integrated seismic measurements at the KTB Exploration Geophysicists.
superdeep drilling site. J. Geophys. Res., 102(B8): 18,267– Helbig, K. & Thomsen, L. 2005. 75-plus years of anisotropy
18,288. in exploration and reservoir seismics: a historical review of
Harris, P.E., Kerner, C. & White, R.E. 1997. Multichannel concepts and methods. Geophysics, 70(6 ): 15.
estimation of frequency-dependent Q from VSP data. Hellweg, M., Spudich, P., Fletcher, J.B. & Baker, L.M. 1995.
Geophys. Prospect., 45: 87–109. Stability of coda Q in the region of Parkfield, California:
Harrison, C.G.A. & Bonatti, E. 1981. The oceanic litho- View from the U.S. Geological Survey Parkfield Dense
sphere. In: The Oceanic Lithosphere. The Sea, Emiliani, C. Seismograph Array. J. Geophys. Res., 100: 2089–2102.
(Ed.). Vol. 7, pp. 21–48. New York: Wiley Herwanger, J.V., Worthington, M.H., Lubbe, R., Binley, A. &
Hasagawa, A., Horiuchi, S. & Umino, N. 1994. Seismic Khazanehdari, J. 2004. A comparison of cross-hole electri-
structure of the northeastern Japan convergent margin: a cal and seismic data in fractured rock. Geophys. Prospect.,
synthesis. J. Geophys. Res., 99(B11): 22,295–22,311. 52: 109–121.
Hasselström, B., Rahm, L. & Scherman, K.A. 1964. Methods Herwanger, J. & Horne, S. 2005. Predicting time-lapse stress
for the determination of the physical and mechanical effects in seismic data. The Leading Edge, December 2005,
properties of rock. Int. Congress on Large Dams, 8. 1234–1242.
Edinburgh. 1(Q 28): 611–625. Paris: ICOLD. Hesler, G.J., Cook, N.G.W. & Myer, L. 1996. Estimation of
Hart, R.S., Anderson, D.L. & Kanamori, M. 1977. The intrinsic and effective elastic properties of cracked media
effect of attenuation on gross earth models. J. Geophys. Res., from seismic testing. 2nd NARMS ’96. Montréal, Québec.
82: 1647–1654. Aubertin, Hassani & Mitri (Eds.). 1: 467–473. Rotterdam:
Hatchell, P. & Bourne, S. 2005. Rocks under strain: strain- Balkema.
induced time-lapse time shifts are observed for depleting Hiramatsu, Y., Hayashi, N. & Furumoto, M. 2000.
reservoirs. The Leading Edge, December 2005, Temporal changes in coda Q1 and b value due to the
1222–1225. static stress change associated with the 1995 Hyogo-ken
Hattema, M.H.H. & de Pater, C.J. 1998. The poro- Nanbu earthquake. J. Geophy. Res., 105: 6141–6151.
mechanical behaviour of Felser sandstone: stress- and Holbrook, W.S., Purdy, G.M., Sheridan, R.E., Glover III, L.,
temperature-dependent. Paper SPE/ISRM 47270, Talwani, M., Ewing, J. & Hutchinson, D. 1994. Seismic
Eurock ’98 Conf., Trondheim, Norway. structure of the U.S. Mid-Atlantic continental margin.
Hauge, P.S. 1981. Measurements of attenuation from verti- J. Geophys. Res., 99(B9): 17,871–17,891.
cal seismic profiles. Geophysics, 37: 1548–1558. Holbrook, W.S., Brocher, T.M., ten Brink, U.S. & Hole, J.A.
Hauksson, E. 2000. Crustal structure and seismicity distri- 1996. Crustal structure of a transform plate boundary:
bution adjacent to the Pacific and North America plate San Francisco Bay and the central California continental
boundary in southern California. J. Geophys. Res., 105: margin. J. Geophys. Res., 101: 22,311– 22,334.
13,875–13,903. Holcomb, D.J. 1988. Cross-hole measurements of velocity and
Hayashi, K. and Saito, H. 2001. Prediction ahead of the tunnel attenuation to detect a disturbed zone in salt at the Waste
face by a high-resolution seismic refraction method with Isolation Pilot Plant. 29th US Symp. on Rock Mechanics,
sources placed in the tunnel. Proc. of 4th Int. Workshop on Minneapolis: MN. Cundall et al. (Eds.). A.A. Balkema,
Application of Geophysics to Rock Engineering, Beijing, China. 633–640.
Int. Soc. Rock Mech. News J., 7(1): 28–33. Hole, J.A., Brocher, T.M., Klemperer, L., Parsons, T.,
Heffer, K.J., Fox, R.J., McGill, C.A. & Koutsabeloulis, N.C. Benz, H.M. & Furlong, K.P. 2000. Three-dimensional
1995. Novel techniques show links between reservoir flow seismic velocity structure of the San Francisco Bay area.
directionality, earth stress, fault structure and geo- J. Geophys. Res., 105(B6 ): 13,859–13,874.
mechanical changes in mature waterfloods. SPE 30711, Holt, R.M., Fjær, E., Furre, A.-K. 1996. Laboratory simula-
SPE Ann. Tech. Conf. Exhib., Dallas: TX. Also SPE J., 2(2): tion of the influence of earth stress change on wave veloci-
91–98, June 1997. ties. In: Seismic Anisotropy, Fjær et al. (Eds.): 180–202.
Heffer, K. 2002. Geomechanical influences in water injec- Society of Exploration Geophysicists.
tion projects: an overview. Oil Gas Sci. Technol. – Rev. IFP, Holt, R.M., Furre, A.-K. & Horsrud, P. 1997. Stress depend-
57(5): 415–422. ent wave velocities in sedimentary rock cores: why and
638 References

why not. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., Hustedt, B. & Clark, R.A. 1999. Source/receiver array direc-
34: 3–4. Paper 128. tivity effects on marine seismic attenuation measurements.
Holt, R.M., Nes, O.-M. & Fjær, E. 2005. In-situ Geophys. Prospect., 47: 1105–1119.
stress dependence of wave velocities in reservoir and over- Hyndman, R.D. 1976. Seismic velocity measurements of
burden rocks. The Leading Edge, December 2005, basement rocks from DSDP leg 37. Initial Rep. Deep Sea
1268–1274. Drill. Proj., 37: 373–387.
Homand-Etienne, F. & Sebaibi, A. 1996. Study of microc- Hyndman, R.D. & Drury, M.J. 1976. The physical proper-
racking of Lac du Bonnet granite. Eurock ’96. Barla (Ed.). ties of oceanic basement rocks from deep drilling on the
2: 1353–1360. Rotterdam: Balkema. Mid-Atlantic ridge. J. Geophys. Res., 81(23): 4042–4052.
Hope, V.S., Clayton, C.R.I. & Barla, G. 1996. Class ‘A’ pre- Hyndman, R.D. 1979. Poisson’s ratio in the oceanic crust – a
dictions of the locations of major rock discontinuities at a review. Tectonophysics, 59: 321–333.
storage cavern site, using seismic tomography. Eurock ’96.
Barla (Ed.). 2: 925–932. Rotterdam: Balkema.
Horn, H.M. & Deere, D.U. 1962. Frictional characteristics I
of minerals. Geotechnique, 12: 319–335.
Horn, S. & MacBeth, C. 1996. The new generation-azimuthal Ikeda, K. 1970. A classification of rock conditions for tun-
variation of amplitude. Extended Abstracts, Society of neling. Proc. of 1st IAEG Congress, Paris. II: 1258–1265. Le
Exploration Geophysicists Workshop, Big Sky, Montana. Comité Français de Géologie de l’Ingénieur.
Horne, S. & MacBeth, C. 1997. AVA observations in Ikeda, K., Ohshima, H. & Sakurai, T. 1981. The property
azimuthal VSPs. 59th Annual Meeting European and the seismic wave velocity of fractured zone. Proc. of
Association Exploration Geophysicists, Expanded Abstracts: Int. Symp. on Weak Rock, Tokyo. 527–532.
P051. Ikeda, R. 1993. In situ stress heterogeneity and crack density
Horne, S., MacBeth, C., Queen, J., Rizer, W. & Cox, V. distribution. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr.,
1997. Fracture characterization from near-set VSP inver- 30(7): 1013–1018. (34th US Symp. on Rock Mechanics,
sion. Geophys. Prospect., 45: 141–164. Madison, WI.)
Horne, S. 2003. Fracture characterization from walkaround Iliev, I.G. 1966. An attempt to estimate the degree of weath-
VSPs. Geophys. Prospect., 51: 493–499. ering of intrusive rocks from their physico-mechanical
Hoshiba, M., Sato, H. & Fehler, M. 1991. Numerical basis of properties. Proc. of 1st ISRM Congress, Lisbon. I: 109–114.
the separation of scattering and intrinsic absorption from full LNEC.
seismogram envelope: a Monte-Carlo simulation of multiple Infanti, N. & Kanji, M.A. 1978. In situ shear strength, nor-
isotropic scattering. Meteorol. Geophys., 42: 65–91. mal and shear stiffness determinations at Agua Vermelha
Hough, S.E., Anderson, J.G., Brune, J., Vernon III, B.F., Project. Proc. of 3rd Int. Cong. of the IAEG, Madrid.
Berger, J., Fletcher, J., Haar, L., Hanks, T. & Baker, L. 1988. Session 2. 2: 172–183.
Attenuation near Anza, California. Bull. Seism. Soc. Am., Infanti, N. & Kanji, M.A. 1990. Estimating the Shear Stiffness
78: 672–691. of rock joints. Rock Joints, Barton, N. & Stephansson, O.
Houtz, R. 1976. Seismic properties of layer 2A in the Pacific. (Eds). Int. Conf., Loen, Norway: 799–804. Rotterdam:
J. Geophys.Res., 81: 6321–6331. Balkema.
Houtz, R. & Ewing, J. 1976. Upper crustal structure as a func- Isacks, B., Oliver, J. & Sykes, L.R. 1968. Seismology and the
tion of plate age. J. Geophys. Res., 81(14 ): 2490–2498. new global tectonics. J. Geophys. Res., 73: 5855–5899.
Hsu, C.-J. & Schoenberg, M. 1993. Elastic waves through a Ishikawa, K., Ochi, H. & Tadaka, S. 1995. Rock mass classi-
simulated fractured medium. Geophysics, 58: 964–977. fication of weathered granite and evaluation of geome-
Hudson, J.A., Jones, E.J.W. & New, B.M. 1980. P-wave chanical parameters for bridge foundation of large scale
velocity measurements in a machine-bored, chalk tunnel. structure in the Honshu-Shikoku Bridge Authority. Rock
Q. J. Engng Geol., 13: 33–43. Northern Ireland: The Foundation. Int. Workshop on Rock Foundation. Tokyo 1995.
Geological Society. 169–176. Rotterdam: Balkema.
Hudson, J.A. 1981. Wave speed and attenuation of elastic ISRM. 1998. Suggested methods for seismic testing within
waves in material containing cracks. Geophys. J. Roy. Astr. and between boreholes. Int. J. Rock Mech. Min. Sci. &
Soc., 64: 133–150. Geomech. Abstr., 25: 447–472.
Hudson, J.A. 1990. Overall elastic properties of isotropic Ito, T. & Zoback, M.D. 2000. Fracture permeability and in
materials with arbitrary distribution of circular cracks. situ stress to 7 km depth in the KTB scientific drillhole.
Geophys. J. Int., 102: 465–469. Geophys. Res. Let., 27: 1045–1048.
Hudson, J.A., Liu, E. & Crampin, S. 1996. The mechanical Ivanovíc, K., Jovanovíc, L., Markovíc, O., Kujundzíc, B. &
properties of materials with interconnected cracks and Radosavljevíc, Z. 1970. Komplexe Untersuchungen mecha-
pores. Geophys. J. Int., 124: 105–112. nischer Charakteristiken der Felsmasse am Profil der
References 639

Bogensperre Mratinje. Proc. of 2nd ISRM Congress, K


Belgrade. III: Theme 8–24. Belgrade: Privredni Pregled.
Iwai, K. 1976. Fundamental studies of fluid flow through a Kabir, M.M.N. & Verschuur, D.J. 2000. A constrained para-
single fracture. PhD Thesis, University of California, metric inversion for velocity analysis based on CFP tech-
Berkeley: CA. 208 pp. nology. Geophysics, 65: 1210–1222.
Iwasaki, T., Yoshii, T., Moriya, T., Kobnayashi, A., Nishiwaki, Kanestrøm, R. & Haugland, K. 1971. The Trans-
M., Tsutsui, T., Iidaka, T., Ikami, A. & Masuda, T. 1994. Scandinavian seismic profile of 1969: profile section 3–4.
Precise P and S wave structures in the Kitakami massif, In: Deep Seismic Sounding in Northern Europe. Vogel, A.
Northern Honshu, Japan, from a seismic refraction exper- (Ed.). Stockholm: Swedish National Science Council,
iment. J. Geophys. Res., 99: 22,187– 22,204. 76–91.
Kang, I.B. & McMechan, G.A. 1994. Separation of intrinsic
and scattering Q based on frequency-dependent ampli-
J tude ratios of transmitted waves. J. Geophys. Res., 99(B12):
23,875–23,885.
Jacobsen, R.S. & Lewis, B.T.R. 1990. The first direct meas- Kappus, M.E., Harding, A.J. & Orcutt, J.A. 1995. A baseline
urements of upper oceanic crustal compressional wave for upper crustal velocity variations along the East Pacific
attenuation. J. Geophys. Res., 95(11): 17,417–17,429. Rise at 13°N. J. Geophys. Res., 100(B4): 6143–6161.
Jaeger, J.C. & Cook, N.G.W. 1977. Fundamentals of Rock Karmis, M., Schilizzi, P. & Agioutantis, Z. 1984. The poten-
Mechanics, Halsted Press. tial and application of sonic wave methods in engineering
Jeng, Y., Tsai, J.-Y. & Chen, S.-H. 1999. An improved rock characterization. Rock Mechanics in Productivity
method of determining near-surface Q. Geophysics, 64(5): and Protection. 25th US Symp. on Rock Mechanics,
1608–1617. Evanston, IL. Dowding & Singh (Eds.). New York: Society
Jin, A., Cao, T. & Aki, K. 1985. Regional change of coda of Mining Engineers, 1083–1091.
Q in the oceanic lithosphere. J. Geophys. Res., 90: Kearey, P. & Brooks, M. 1994. An Introduction to Geophysical
8651–8659. Exploration. Geoscience Texts, Vol. 4, Oxford: Blackwell
Jin, A. & Aki, J. 1986. Temporal change in coda Q before the Scientific Publications.
Tangsham earthquake of 1976 and the Haicheng earth- Kearey, P. & Vine, F.J. 1996. Global Tectonics, 2nd edn. 333
quake of 1975. J. Geophys. Res., 91: 665–673. pp. Oxford: Blackwell Science.
Jizba, D.L. 1991. Mechanical and acoustical properties of sand- Keehn, N.A. & Kanasewich, E.R. 1987. Attenuation from
stones and shales. PhD Dissertation, Stanford University. VSP data collected on Melville Island. Can. J. Soc. Explor.
Johansen, T.A., Jakobsen, M., Agersborg, R. & Ruud, B.O. Geophys., 23(1): 73–84.
2004. The P-P and P-S response of a fractured reservoir. Kelemen, P.B. & Holbrook, W.S. 1995. Origin of thick, high-
SEG Int. Exposition and 74th Annual Meeting, Denver, Co. velocity igneous crust along the U.S. East Coast Margin.
Johnston, D.H., Toksöz, M.N. & Timur, A. 1979. J. Geophys. Res., 100(B7 ): 10,077–10,094.
Attenuation of seismic waves in dry and saturated rocks: Kelsall, P.C., Watters, R.J. & Franzone, J.G. 1986. Engineering
II. Mechanisms. Geophysics, 44(4 ): 691–711. characterization of fissured, weathered dolerite and vesicular
Johnston, D.H. 1986. VSP detection of fracture-induced basalt. Rock Mechanics: Key to Energy Production. 27th
velocity anisotropy. 56th Annual Int. Meeting, Society of US Symp. on Rock Mechanics, Tuscaloosa: AL. Hartman
Exploration Geophysicists, Houston: TX. (Ed.): 77–84. New York: Society of Mining Engineers.
Johnston, J.E. & Christensen, N.I. 1995. Seismic anisotropy Kennett, B.L.N. & Orcutt, J.A. 1976. A comparison of travel
of shales. J. Geophys. Res., 100: 5991–6003. time inversions for marine refraction profiles. J. Geophys.
Jones, T. & Nur, A. 1983. Velocity and attenuation in sand- Res., 81(23): 4061–4070.
stones at elevated temperatures and pressures. Geophys. Khazanehdari, J., McCann, C., Sothcott, J. & Astin, T.R.
Res. Lett., 10: 140–143. 1998. The inter-relationships between velocity and attenu-
Jones, T.D. 1986. Pore fluids and frequency-dependent wave ation from VSP survey. Technical Proc. of 60th annual meet-
propagation in a vertical transversely isotropic medium. ing of the European Association of Geoscientists and Engineers,
Field experiment and model study. Geophysics, 51: Leipzig. Paper No. 10–28.
1939–1953. Kikuchi, K., Saito, K. & Kusunoki, K.-I. 1982. Geotechnically
Jongmans, D. & Malin, P.E. 1995. Vertical profiling of integrated evaluation on the stability of dam foundation
microearthquake S waves in the Varian well at Parkfield, rocks. Int. Congress on Large Dams, 14. Rio de Janeiro
California. Bull. Seism. Soc. Am., 85: 1805–1820. 1982. 2(Q53): 49–73. Paris: ICOLD.
Juhlin, C. 1995. Imaging of fracture zones in the Finnsjönn King, M.S. 1966. Wave velocities in rocks as a function of
area, central Sweden, using the seismic reflection method. changes in overburden pressure and pore fluid saturants.
Geophysics, 60(1): 66–75. Geophysics, 31: 50–73.
640 References

King, M.S., Pandit, B.I. & Stauffer, M.R. 1978. Quality of Knipe, R.J., Jones, G. & Fisher, Q.J. (1998). Faulting, fault seal-
rock masses by acoustic borehole logging. Proc. of 3rd ing and fluid flow in hydrocarbon reservoirs: an introduction.
IAEG Congress, Madrid. IV(1): 156–164. Madrid: Imprime In: Faulting, Fault Sealing and Fluid Flow in Hydrocarbon
ADOSA. Reservoirs, Jones, G., Fisher, Q.J. & Knipe, Q.J. (Eds.).
King, M.S., Myer, L.R. & Rezowalli, J.J. 1984. Cross-hole Geological Society London Special Publication 147: i–xxi.
acoustic measurements in basalt. Rock Mechanics in Pro- Knopoff, L. 1964. ‘Q’. Rev. Geophys., 2(4): 645–654.
ductivity and Protection. 25th US Symp. on Rock Mechanics, Publication 391, Institute of Geophysics and Planetary
Evanston, IL. Dowding & Singh (Eds.). New York: Society Physics, UCLA.
of Mining Engineers, 1053–1061. Kranz, R.L., Frankel, A.D., Engelder, T. & Scholtz, C.H.
King, M.S., Myer, L.R. & Rezowalli, J.J. 1986. Experimental 1980. The permeability of whole and jointed Barre gran-
studies of elastic-wave propagation in a columnar-jointed ite. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 16:
rock mass. Geophys. Prospect., 34: 1185–1199. 225–234: Erratum, 17: 237–238.
King, M.S., Chaudhry, N.A. & Shakeel, A. 1995. Experimental Kujundzíc, B. & Grujíc, N. 1966. Correlation between static
ultrasonic velocities and permeability for sandstones with and dynamic investigations of rock mass “in situ”. Proc. of
aligned cracks. Int. J. Rock Mech. Min. Sci. & Geomech. 1st ISRM Congress, Lisbon. 1: 565–570. LNEC.
Abstr., 32: 155–163. Kujundzíc, B., Jovanovic, L. & Radosavljevic, Z. 1970.
King, M.S., Shakeel, A. & Chaudhry, N.A., 1997. Acoustic Solution du revêtement d’une galerie en charge par applica-
wave propagation and permeability in sandstones with sys- tion d’injections à haute pression. Proc. of 2nd ISRM
tems of aligned cracks. In: Developments in Petrophysics, Congress, Belgrade. II: Theme 4–66. Belgrade: Privredni
Lovell, M.A. & Harvey, P.K. (Eds.). Geological Society of Pregled.
London, Special Publication No. 122, 69–85. Kujundzíc, B. 1979. Use of tests and monitoring in the
King, M.S. 2002. Elastic wave propagation in and permea- design and construction of rock structures. Proc. of 4th Int.
bility for rocks with multiple parallel fractures. Technical Congr. Rock Mech. (ISRM), Montreux: 3: 181–186.
Note, Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 39: Rotterdam: Balkema.
1033–1043. Kuster, G.T. & Toksöz, M.N. 1974. Velocity and attenuation
King, M.S. & Marsden, J.R. 2002. Velocity dispersion of seismic waves in two-phase media: Part II, Experimental
between ultrasonic and seismic frequencies in brine- results. Geophysics, 39: 587–606.
saturated reservoir sandstones. Geophysics, 67: 254–258. Kvamme, L.B. & Havskov, J. 1989. Q in Southern Norway.
King, M.S. 2005. Rock-physics developments in seismic explo- Bull. Seism. Soc. Am., 79: 1575–1588.
ration: a personal 50-year perspective. Geophysics, 70(6 ): 3–8.
Kinoshita, S. 1994. Frequency-dependent attenuation of
shear waves in the crust of the Southern Kanton Area, L
Japan. Bull. Seism. Soc Am., 84: 1387–1396.
Kirkpatrick, R.J. 1979. The physical state of the oceanic Laine, R.G.T. 1964. Rock foundations. Diagnosis of
crust: results of downhole geophysical logging in the Mid- mechanical properties and treatment. Proc. of 8th Int.
Atlantic ridge at 23°N. J. Geophys. Res., 84(B1): 178–188. Congress on Large Dams, Edinburgh, 1964. 1: 141–165.
Kishida, K. 1999. Experimental study on the estimation of Paris: ICOLD.
discontinuity characteristics and mechanical properties of Landrø, M. & Stammeijer, J. 2004. Quantitative estimation
jointed rock masses. PhD Thesis, Civil Engineering of compaction and velocity changes using 4D impedance
Department, University of Kyoto, Japan, 203 pp. and traveltime changes. Geophysics, 69(4 ): 949–957.
Kjartansson, E. 1977. Constant Q: wave propagation and Lashkaripour, G.R. & Passaris, E.K.S. 1995. Correlations
attenuation. EOS, Trans. AGU., 58: 1183. between index parameters and mechanical properties of
Kjartansson, E. 1979. Constant Q-wave propagation and shales. Proc. of 8th ISRM Congress, Tokyo. Fuji (Ed.). 1:
attenuation. J. Geophys. Res., 84(B9): 4737–4748. 257–261. Rotterdam: Balkema.
Klimentos, T. & McCann, C. 1990. Relationships among com- Laubach, S., Marrett, R. & Olson, J. 2000. New directions
pressional wave attenuation, porosity, clay content and in fracture characterization. The Leading Edge, July 2000,
permeability in sandstones. Geophysics, 55: 998–1014. 704–711.
Klimentos, T. 1991. The effects of porosity-permeability-clay Laubach, S.E., Gale, J.F.W. & Olson, J.E. 2002. Are open
content on the velocity of compressional waves. Geophysics, fractures necessarily aligned with maximum horizontal
56: 1930–1939. stress? American Association of Petroleum Geologists. (AAPG:
Klimentos, T. 1995. Attenuation of P- and S-waves as a Fractured Reservoirs and Multi-Phase Fluid Flow), Annual
method of distinguishing gas and condensate from oil and Meeting, Houston, TX.
water. Geophysics, 60: 447–458. Laubach, S.E. 2003. Practical approaches to identifying
Knight, R., Dvorkin, J. & Nur, A. 1998. Acoustic signatures sealed and open fractures. AAPG (American Association of
of partial saturation. Geophysics, 60: 132–138. Petroleum Geologists). Bull., 87(4): 561–579.
References 641

Laurent, J., Boutéca, M.J., Sarda, J.-P. & Bary, D. 1995. Pore Liu, E., Maultzsch, S., Chapman, M. & Li, X.Y. 2003b.
pressure influence in the poroelastic behaviour of rocks: Frequency-dependent seismic anisotropy and its implica-
experimental studies and results. SPE Formation Evaluation, tion for estimating fracture size in low porosity reservoirs.
June 1993, 117–122. The Leading Edge, July 2003: 662–665.
Leary, P.C. & Henyey, T.L. 1985. Anisotropy and fracture Liu, E., Vlastos, S. & Li, X.Y. 2004. Modeling seismic wave
zones about a geothermal well from P-wave velocity pro- propagation during fluid injection in a fractured network:
files. Geophysics, 50(1): 25–36. effects of pore fluid pressure on time-lapse seismic signa-
Leary, P.C., Crampin, S. & McEvilly, T.V. 1990. Seismic tures. The Leading Edge, August 2004: 778–783.
fracture anisotropy in the earth’s crust: an overview. Liu, E. 2005. Effects of fracture aperture and roughness on
J. of Geophys. Res., 95(B7): 11,105–11,114. hydraulic and mechanical properties of rocks: implication of
Leary, P.C. & Abercrombie, R.E. 1994. Frequency- seismic characterization of fractured reservoirs. J. Geophys.
dependent crustal scattering and absorption at 5–160 Hz Eng., 2: 38–47, Nanking Inst. of Geophys. Prospect.
from Coda decay observed at 2.5 km depth. Geophys. Res. Liu, Y., Booth, D.C., Crampin, S., Evans, R. & Leary, P.
Lett., 21: 971–974. 1993. Shear-wave polarizations and possible temporal vari-
Lecomte, I., Gjøystdal, H., Dahle, A. & Pedersen, O.C. ations in shear-wave splitting at Parkfield. Can. J. Explor.
2000. Improving modelling and inversion in refraction Geophys., 29: 380–390.
seismics with a first-order Eikonal solver. Geophys. Liu, Y., Crampin, S. & Abercrombie, R.E. 1997. Shear-wave
Prospect., 48: 437–454. anisotropy and the stress field from borehole recordings at
Lee, H. & Haimson, B.C. 2006. Borehole breakouts and 2.5 km depth at Cajon Pass. Geophys. J. Int., 129: 439–449.
in situ stress in sandstone. In situ Rock Stress. Proc. of Int. Liu, Y., Teng, T.-L., & Ben-Zion, Y. 2005. Near-surface
Conf., Lu, Li, Kjorholt & Dahle (Eds.): 201–207. seismic anisotropy, attenuation and dispersion in the
London: Taylor & Francis. aftershock region of the 1999 Chi-Chi earthquake.
Li, X., Cui, L. & Rogiers, J.-C. 1998. Thermo-poroelastic Geophys. J. Int., 160: 695–706.
analyses of inclined boreholes. Paper SPE/ISRM 47296, Liu, Y.-Q. 2003. Analysis of P-wave seismic reflection data for
Eurock 98 Conf., Trondheim, Norway. azimuthal anisotropy. PhD Thesis, University of Edinburgh.
Liao, Q. & McMechan, G.A. 1997. Tomographic imaging of Londe, P. 1979. Panel discussion: dynamic testing. Int. Symp.
velocity and Q, with application to crosswell seismic data on The Geotechnics of Structurally Complex Formations. Capri
from the Gypsy Pilot Site, Oklahoma. Geophysics, 62(6 ): 1977. II: 279–281. Associazione Geotecnica Italiana.
1804–1811. Louis, C. 1969. A study of groundwater flow in jointed rock and
Link, H. 1964. Evaluation of elasticity moduli of dam foun- its influence on the stability of rock masses. Rock Mechanics
dation rock determined seismically in comparison to those Res. Report, 10, Imperial College. (Translation of PhD
arrived at statically. Int. Congress on Large Dams, 8. Thesis).
Edinburgh 1964. 1(Q28): 833–858. Paris: ICOLD. Lubbe, R. 2005. A field and laboratory investigation of the
Liu, E., Crampin, S., Queen, J.H. & Rizer, W.D. 1993. compliance of fractured rock. PhD Thesis, Oxford
Velocity and attenuation anisotropy caused by microc- University.
racks and macrofractures in a multiazimuth reverse VSP. Lubbe, R. & Worthington, M. 2006. A field investigation of
Can. J. Explor. Geophys., 29: 177–188. fracture compliance. Geophys. Prospect., 54(3): 319–331.
Liu, E. & Li, X.-Y. 1998. Estimation of pore or fracture Lubbe, R., Sothcott, J., Worthington, M.H., McCann, C. &
fluids from seismic measurements. 68th Annual Int. SEG Astin, T. 2006. Laboratory estimates of fracture compliance.
Meeting, New Orleans, 1624–1627. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr. (In press).
Liu, E., Hudson, J.A. & Pointer, T. 2000. Equivalent Lucet, N. 1989. Vitesse et Attenuation des ondes élastiques
medium representation of fractured rock. J. Geophys. Res., soniques et ultrasoniques dans les roches sous pression de
105(B2): 2981–3000. confinement. PhD Dissertation, University of Paris.
Liu, E., Li, X.Y. & Queen, J.H. 2000. Discrimination of pore Lucet, N. & Zinszner, B. 1992. Effects of heterogeneities and
fluids from P and converted shear-wave AVO analysis. anisotropy on sonic and ultrasonic attenuation in rocks.
Anisotropy 2000: Fractures, Converted Waves and Case Studies. Geophysics, 57(8 ): 1018–1026.
Tulsa, OK: Society of Exploraion Geophysicists. Ludwig, J.W., Nafe, J.E. & Drake, C.L. 1970. Seismic
Liu, E., and Li, X.-Y. 2001. Seismic detection of fluid satur- refraction. In: The Sea. Maxwell, A.E. (Ed.). 4: 53–84.
ation in aligned fractures. 9th Int. Workshop on Seismic New York: Wiley.
Anisotropy. Geophysics, 66(4 ): Abstract. Lumley, D.E. 2001. Time-lapse seismic reservoir monitor-
Liu, E., Queen, J.H., Li, X.Y., Chapman, M., Maultzsch, S., ing. Geophysics, 66: 50–53.
Lynn, H.B. & Chesnokov, E.M. 2003a. Observation Lykoshin, A.G., Yaschenko, S.G., Mikhailov, A.D.,
and analysis of frequency-dependent anisotropy from a Savitch, A.I. & Koptev, V.J. 1971. Investigation of rock
multicomponent VSP at Bluebell-Altamont Field, Utah. jointing by seismo-acoustic methods. Int. Symp. on
J. Appl. Geophys., 54: 319–333. Fissuration des Roches, Nancy, France, I–19.
642 References

Lynn, H.B. & Thomson, L.A. 1986. Reflection shear wave data Majer, E.L., Peterson, J.E., Daley, T., Kaelin, B., Myer, L.,
along the principle axes of azimuthal anisotropy. 56th Annual Queen, J., D’Onfro, P. & Rizer, W. 1997. Fracture detec-
Int. Meeting SEG. Tulsa, OK: Society of Exploration tion using crosswell and single well surveys. Geophysics,
Geophysicists. 62(2): 495–504.
Lynn, H.B., Simon, K.M., Bates, C.R., Layman, M., Makurat, A., Barton, N., Chryssanthakis, P. & Monsen, K.
Schneider, R. & Jones, M. 1995. Use of anisotropy in 1990. Jointed rock mass modelling. Rock Joints, Barton, N. &
P-wave and S-wave data for fracture characterization in a Stephansson, O. (Eds.). Int. Conf., Loen, Norway. 647–
naturally fractured gas reservoir. The Leading Edge, August 656. Rotterdam: Balkema.
1995: 887–893. Makurat, A., Barton, N., Rad, N.S. & Bandis, S. 1990.
Lynn, H.B., Beckham, W.E., Simon, K.M., Bates, C.R., Joint conductivity variation due to normal and shear
Layman, M. & Jones, M. 1999a. P-wave and S-wave deformation. Rock Joints, Barton, N. & Stephansson, O.
azimuthal anisotropy at a naturally fractured gas reservoir, (Eds.). Int. Conf., Loen, Norway. 535–540. Rotterdam:
Bluebell-Altamont Field, Utah. Geophysics, 64: 1312–1328. Balkema.
Lynn, H.B., Campagna, D., Simon, K.M. & Beckham, W.E. Makurat, A., Barton, N., Tunbridge L. & Vik, G. 1990. The
1999b. Relationship of P-wave seismic attributes, azimu- measurement of the mechanical and hydraulic properties of
thal anisotropy, and commercial gas pay in 3-D P-wave rock joints at different scales in the Stripa Project. Rock Joints,
multiazimuth data, Rulison Field, Piceance Basin, Colorado. Barton, N. & Stephansson, O. (Eds.). Int. Conf., Loen,
Geophysics, 64(4 ): 1293–1311. Norway. 541–548. Rotterdam: Balkema.
Lynn, H.B., Weathers, L. & Beckham, W. 2000. The Mobil Makurat, A. 1996. Fracture flow and fracture cross flow as a
Onshore Texas 3D full-azimuth full-offset P-wave survey. function normal stress and shear displacement, Extended
9th Int. Workshop on Seismic Anisotropy. Extended Abstract. Abstract: 10th Kongsberg Seminar on Fractures, Fluid Flow
Lynn, H.B. 2004. The winds of change: anisotropic rocks – and Transport in Fractures, Kongsberg, Norway.
their preferred direction of fluid flow and their associated Malin, P.E., Walker, J.A., Borcherdt, R.D., Cranswick, E.,
seismic signatures. Part 1: The Leading Edge, November Jensen, E.G. & Van Schaak, J. 1988. Vertical seismic
2004, 23(11): 1156–1162, Part 2: The Leading Edge, profiling of Oroville microearthquakes: velocity spectra and
December 2004, 23(12): 1258–1268. particle motion as a function of depth. Bull. Seism. Soc. Am.,
78: 401–420.
Mandal, P., Padhy, S., Rastogi, B.K., Satyanarayana, H.V.S.,
Kousalya, M., Vijayraghavan, R. & Srinivasan, A. 2001.
M Aftershock activity and frequency-dependent low coda Q c
in the epicentral region of the 1999 Chamoli Earthquake of
MacBeth, C. 1999. Azimuthal variation in P-wave signatures Mw 6.4. Pure Appl. Geophys., 158: 1719–1735.
due to fluid flow. Geophysics, 64: 1181–1192. Mandal, P., Jainendra, T.S., Joshi, S., Kumar, S., Bhunia, R.,
MacBeth, C. & Li, X.Y. 1999. AVD – An emerging new Rastogi, B.K. 2004. Low coda Q c in the epicentral region
marine technology for reservoir characterization: acquisition of the 2001 Bhuj Earthquake of Mw 7.7. Pure Appl.
and application. Geophysics, 64: 1153–1159. Geophys., 161: 1635–1654.
MacBeth, C. 2004. A classification for the pressure-sensitivity Marmureanu, G., Bratosin, D. & Cioflan, C.O. 2000. The
properties of a sandstone rock frame. Geophysics, 69(2): dependence of Q with seismic-induced strains and fre-
497–510. quencies for surface layers from resonant columns. Pure
MacDonal, F.J., Angona, F.A., Mills, R.L., Sengbush, R.L., Van Appl. Geophys., 157: 269–279.
Nostrand, R.G. & White, J.E. 1958. Attenuation of shear Marson-Pidgeon, K. & Savage, M.K. 1997. Frequency-
and compressional waves in Pierre shale. Geophysics, 23: dependent anisotropy in Wellington, New Zealand.
421–439. Geophys. Res. Lett., 24(24): 3297–3300.
Macdonald, K.C. 1982. Mid-Ocean ridges: fine scale tectonic, Martel, S.J. & Peterson, Jr., J.E. 1991. Interdisciplinary char-
volcanic and hydrothermal processes within the plate acterization of fracture systems at the US/BK site, Grimsel
boundary zone. Annu. Rev. Earth Planet. Sci., 10: 155–190. Laboratory, Switzerland. Int. J. Rock Mech. Min. Sci. &
Maia, A.C., Poiate, E.J., Falčao, J.L. & Coelho, L.F.M. 2005. Geomech. Abstr., 28(4): 295–323.
Triaxial creep tests in salt applied in drilling through thick salt Martin, C.D., Davison, C.C. & Kozak, E.T. 1990. Charac-
layers in Campos Basin, Brazil. ISRM News J., 9(1): 14–24. terizing normal stiffness and hydraulic conductivity of a
Maini, T. & Hocking, G. 1977. An examination of the feasi- major shear zone in granite. Rock Joints, Barton, N. &
bility of the hydrologic isolation of a high level waste Stephansson, O. (Eds.). Int. Conf., Loen, Norway. 549–556.
repository in crystalline rock. Geologic disposal of high- Rotterdam: Balkema.
level radioactive waste session, Annual Meeting of the Geol. Martin, C.D., Young, R.P. & Collins, D.S. 1995. Monitoring
Soc. of America, Seattle, Washington. progressive failure around a tunnel in massive granite. Proc. of
References 643

8th Int. Conf. of Int. Soc. Rock Mech., Tokyo, Japan. Fuji, T. Mavko, G.M. & Jizba, D. 1991. Estimating grain-scale fluid
(Ed.). 627–634. Rotterdam: Balkema. effects on velocity dispersion in rocks. Geophysics, 56:
Martin, C.D., Read, R.S. & Martino, J.B. 1997. Observation 1940–1949.
of brittle failure around a circular test tunnel. Int. J. Rock Mavko, G., Mukerji, T. & Dvorkin, J. 1998. The Rock
Mech. Min. Sci. & Geomech. Abstr., 34: 1065–1073, Elsevier. Physics Handbook: Tools for Seismic Analysis in Porous
Masuda, H. 1964. Utilisation of elastic longitudinal wave Media, Cambridge University Press.
velocity for determining the elastic property of dam foun- Mavko, G. 2002. Rock physics signatures of reservoir
dation rocks. Int. Congress on Large Dams, 8. Edinburgh heterogeneities. American Association of Petroleum
1964. 1(Q28): 253–269. Paris: ICOLD. Geologists. (AAPG Rock Physics: The Missing Link
Mathisen, M.E., Vasilou, A.A., Cunningham, P., Shaw, J., between Geology, Geophysics and Production), Annual
Justice, J.H. & Guinzy, N.J. 1995. Geophysics, 60(3): Meeting, Houston: TX.
631–650. Maxwell, S.C. & Young, R.P. 1993. Seismic imaging of blast
Matthews, M.C., Hope, V.S. & Clayton, C.R.I. 1997. The damage. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr.,
geotechnical value of ground stiffness determined using 30(7 ): 1435–1440.
seismic methods. Modern Geophysics in Engineering Maxwell, S.C. & Young, R.P. 1996. Seismic imaging of rock
Geology, McCann, Eddleston, Fenning & Reeves (Eds.). mass responses to excavation. Int. J. Rock Mech. Min.
113–123. Geological Society Engineering Geology Special Sci. & Geomech. Abstr., 33(7 ): 713–724.
Publication. Mayeda, K., Koyanagi, S., Hoshiba, M., Aki, K. & Zeng, Y.
Maultzsch, S., Liu, E. & Chapman, M. 2002. Modelling 1992. A comparative study of scattering, intrinsic and
attenuation anisotropy and its interpretation in terms of coda Q1 for Hawaii, Long Valley, and Central California
fracture properties. Abstract 73rd Annual Int. SEG between 1.5 and 15.0 Hz. J. Geophys. Res., 97:
Meeting, Dallas: TX. 6643–6659.
Maultzsch, S., Chapman, M., Liu, E. & Li, X.Y. 2003. McCaffree Pellerin, C.L. & Christensen, N.I. 1998.
Modelling frequency-dependent seismic anisotropy in Interpretation of crustal seismic velocities in the San Gabrial-
fluid–saturated rock with aligned fractures: implication of Mojave region, Southern California. Tectonophysics, 286:
fracture size estimaton from anisotropic measurements. 253–271.
Geophys. Prospect., 51: 381–392. McCann, D.M., Grainger, P. & McCann, C. 1975. Inter-
Maultzsch, S., Chapman, M., Liu, E. & Li, X.Y. 2005. borehole acoustic measurements and their use in engineer-
Observation and modelling of anisotropic attenuation in ing geology. Geophys. Prospect., 23: 50–69.
VSP data. Seismic Q–observation, mechanisms, and inter- McDowell, P.W., Barla, G. & Sharp, J.C. 1992. Investigation
pretation. Workshop at 67th EAGE Conference, Madrid. of rock mass characteristics by geophysical methods for an
Extended Abstract: Q-08, Convenors: Hargreaves, Best underground storage project site in weak jointed chalk.
and Lancaster. ISRM Symp. on Eurock’92. Chester, UK. 169–173. London:
Maury, V. 1987. Observations, recherches et résultants récents Thomas Telford.
sur la mécanismes de ruptures autour de galleries isolées. McDowell, P.W. 1993. Seismic investigation for rock engin-
Proc. of 6th Int. Congr. Rock Mech. (ISRM), Montreal, eering. Comprehensive Rock Engineering, Hudson, J.A.
Herget, G. & Vongpaisal (Eds.): 2: 1119–1128. (Editor in Chief ). Vol. 3. Rock Testing and Site
Rotterdam: Balkema. Characterization. 619–634. Oxford: Pergamon Press.
Maury, V. & Guenot, A. 1995. Practical advantages of mud McKenzie, C.K., Stacey, G.P. & Gladwin, M.T. 1982. Ultra-
cooling systems for drilling. SPE Drillings & Completions, sonic characteristics of a rock mass. Int. J. Rock Mech. Min.
10(1), 42–48. Sci. & Geomech. Abstr., 19: 25–30.
Maury, V., Piau, J.-M. & Hallé, G. 1996. Subsidence induced Miller, D.E., Leaney, S. & Borland, W.H. 1994. An in situ
by water injection in water sensitive reservoir rocks: the estimation of anisotropic elastic moduli for a submarine
example of Ekofisk. Paper SPE 36890, SPE European shale. J. Geophys. Res., 99(B11): 21,659–21,665.
Petroleum Conf., Milan, Italy. Miller, R.P. 1965. Engineering classification and index
Mavko, G. & Nur, A. 1975. Melt squirt in the asthenosphere. properties for intact rock. PhD Thesis, University of
J. Geophys. Res., 80: 1444–1448. Illinois, 332 pp.
Mavko, G. 1979. Frictional attenuation: an inherent ampli- Mitani, S., Iwai, T. & Isahai, H. 1987. Relations between con-
tude dependence. J. Geophys. Res., 84: 4769–4776. ditions of rock mass and TBM’s feasibility. Proc. of 6th Int.
Mavko, G.M. & Nur, A. 1979. Wave attenuation in partially Congr. Rock Mech. (ISRM), Montreal, Herget, G. &
saturated rocks. Geophysics, 44: 161–178. Vongpaisal (Eds.): 1: 701–704. Rotterdam: Balkema.
Mavko, G.M., Kjartansson, E. & Winkler, K. 1979. Seismic Mitchell, B.J. & Hwang, H.J. 1987. Effect of low Q sedi-
wave attenuation in rocks. Rev. of Geophys. Space Phys., 17: ments and crustal Q on Lg attenuation in the United States.
1155–1164. Bull. Seism. Soc. Am., 77(4): 1197–1210.
644 References

Mitchell, B.J. 1995. Anelastic structure and evolution of the Proc. of SPE European Petroleum Conf., Paris, SPE Paper
continental crust and upper mantle from seismic surface No. 65180. Tulsa, OK: Society of Petroleum Engineers.
wave attenuation. Rev. Geophys., 33: 441–462. New, B.M. & West, G. 1980. The transmission of compres-
Mitchell, B.J., Pan, Y., Xie, J. & Cong, L. 1997. Lg coda Q sional waves in jointed rock. Eng. Geol., 15: 151–161.
variation across Eurasia and its relation to crustal evolution. New, B.M. 1985. An example of tomographic and Fourier
J. Geophys. Res., 102: 22,767–22,779. microcomputer processing of seismic records. Q. J. Eng.
Moen, P.A. 2004. The greener side of grouting in Norway, Geol., 18: 335–344. Northern Ireland: The Geological
Tunnels and Tunnelling International, June 2004: 14–16. Society.
Mohorovičič, A. 1909. Das bebenvom 8, X, Jahrb. Meterol. Newmark, R.L., Anderson, R.N., Moos, D. & Zoback, M.D.
Obs. Zagreb, 9: 1–63. 1985. Structure, porosity and stress regime of the upper
Molnar, P. & Oliver, J. 1969. Lateral variations of attenu- oceanic crust: sonic and ultrasonic logging of DSDP
ation in the upper mantle and discontinuities in the litho- Hole 504B. Tectonophysics, 118: 1–42.
sphere. J. Geophys. Res., 74: 2648–2682. Nihei, K.T. & Cook, G.W. 1992. Seismic behavior of frac-
Molyneux, J.B. & Schmitt, D.R. 2000. Compressional-wave tures under normal compression. 33rd US Symp. on Rock
velocities in attenuating media: a laboratory physical model Mechanics: Rock mechanics. Tillerson & Wawersik (Eds.).
study. Geophysics, 65(4 ): 1162–1167. 989–997. Rotterdam: Balkema.
Monsen, K. 2001. Acoustic velocity in fractured rock. Niini, H. & Manunen, T. 1970. Seismic sounding as indicator
J. Geophys. Res., 106(B7 ): 13,261–13,267. of engineering-geologic properties of bedrock in Finland.
Mooney, W.D., Laske, G. & Masters, T.G. 1998. CRUST Proc. of 1st IAEG Congress, Paris, II: 753–761. Le Comité
5.1: a global crustal model at 5°  5°. J. Geophys Res., Français de Géologie de l’Ingénieur.
103(B1): 727–747. Nilsen, B. 1997. Hvorfor stemmer ikke terrenget med kartet –
Moos, D. & Zoback, M.D. 1983. In-situ studies of velocity er ikke de ingeniørgeologiske forundersøkelsene gode nok?
in fractured crystalline rocks. J. Geophys. Res., 88(B3): Cases of unexpected ground conditions – how much are
2345–2358. they related to insufficient engineering geological investiga-
Mouraz Miranda, A. & Mello Mendes, F. 1987. Rock weath- tions? Fjellsprengningsteknikk – bergmekanikk – geoteknikk,
ering as a drillability parameter, Proc. of 6th Int. Congr. 1997. 5.1–5.10. Oslo: Tapir.
Rock. Mech. (ISRM), Montreal, Herget, G. & Vongpaisal Nilsen, B. 1998. Subsea rock tunnels – some Norwegian experi-
(Eds.): 1: 699–700. Rotterdam: Balkema. ences in geo-investigation and construction. Design and
Myer, L.R., Hopkins, D., Peterson, J.E. & Cook, N.G.W. Construction in Mining Petroleum and Civil Engineering,
1995. Seismic wave propagation across multiple fractures. Ayres da Silva, Quadros, Gonçalves, (Eds.). Brazil: São Paulo.
Int. Conf. on Fractured and Jointed Rock Masses, Lake Nishioka, G.K. & Michael, A.J. 1990. A detailed study of
Tahoe: CA. Myer, Tsang, Cook & Goodman (Eds.). the seismicity of the Middle Mountain zone at Parkfield,
105–109. Rotterdam: Balkema. California: temporal variations as possible precursors.
J. Geophys. Res., 193: 577–588.
Nishioka, K. & Aoki, K. 1998. Rapid tunnel excavation by
N hard rock TBMs in urban areas. In: Tunnels and Metropolises,
Negro Jr. & Ferreira, (Eds.). 655–661. Rotterdam: Balkema.
Nasu, N. 1940. Studies on the propagation of an artificial Nolet, G. (Ed.). 1987. Seismic Tomography with Applications
earthquake wave through superficial soil or sand layers, and in Global Seismology and Exploration Geophysics., D. Reidel
the elasticity of soil and sand. [in Japanese]. Bull. Earthq. Res. Publishing Co. (Springer).
Inst., 18: 289–304. Japan: Tokyo Imperial University. Nord, G., Olsson, P. & By, T.L. 1991. Geophysical ground
Navalón, N., Gaztanaga, J.M. & Lopez Marinas, J.M. 1987. probing in TBM Tunnelling. Proc. of 7th ISRM Congress,
Cortes-La Muela hydroelectric project. The foundations of Aachen. Wittke (Ed.). 1: 581–586. Rotterdam: Balkema.
Cortes and El Naranjero dams. Proc. of 6th Int. Congr. Rock Norton, D. & Knapp, R. 1977. Transport phenomena in
Mech. (ISRM), Montreal, Herget, G. & Vongpaisal (Eds.): hydrothermal systems: the nature of porosity. Am. J. Sci.,
1: 449–455. Rotterdam: Balkema. 277: 913–936.
Nehlig, P. & Juteau, T. 1988. Flow porosities, permeabilities Nunn, K.R., Barker, R.D. & Bamford, D. 1983. In situ seis-
and preliminary data on fluid inclusions and fossil thermal mic and electrical measurements of fracture anisotropy in
gradients in the crustal sequence of the Sumail ophiolite the Lincolnshire Chalk. Q. J. Eng. Geol., 16: 187–195.
(Oman). Tectonophysics, 151: 199–221. Northern Ireland: The Geological Society.
Nelson, R.A. 2001. Geologic Analysis of Naturally Fractured Nur, A. & Simmons, G. 1969. The effect of saturation on veloc-
Reservoirs. Boston: Gulf Professional Publishing. ity in low porosity rocks. Earth Planet. Sci. Lett., 7: 183–193.
Nes, O.M., Holt, R.M. & Fjær, E. 2000. The reliability of Nur, A. 1971. Effects of stress on velocity anisotropy in rocks
core data as input to seismic reservoir monitoring studies. with cracks. J. Geoph. Res., 76(8): 2022–2034.
References 645

Nur, A. 1974. Tectonophysics: the study of relations between Olsson, R. & Barton, N. 2001. An improved model for hydro-
deformation and forces in the earth. Proc. of 4th Int. Cong. mechanical coupling during shearing of rock joints. Int.
on Rock Mechanics, Denver, CO, 1A: 243–317. J. Rock Mech. Min. Sci. & Geomech. Abstr., 38: 317–329.
Nur, A. 1987. Seismic rock properties for reservoir descrip- Pergamon.
tions and monitoring (Chapter 9). In: Seismic Tomography, Onedera, T.F. 1963. Dynamic investigation of foundation
Nolet, G. (Ed.). 203–237. D. Reidel Publishing Co. rocks in situ. 5th US Symp. on Rock Mechanics, Minnesota.
Nur, A. 1989. Four-dimensional seismology and (true) direct 517–533. New York: Pergamon.
detection of hydrocarbons: the petrophysical basis. The
Leading Edge, August 1989, 8: 30–36.
Nur, A., Marion, D. & Yin, H. 1991. Wave velocities in sedi- P
ments. In: shear Waves in Marine Sediments, Hovem, J.M.
Richardson, M.D. & Stoll, R.D. (Eds.). 131–140. Paffenholz, J. & Burkhardt, H. 1989. Absorption and modu-
Dordrecht: Kluwer Academic Publishers. lus measurements in the seismic frequency and strain
Nur, A.M. & Wang, Z. (Eds.). 1992. Seismic and Acoustic range on partially saturated sedimentary rocks. J. Geophys.
Velocities in Reservoir Rocks. Vol. 1–2. Geophysical reprint Res., 94: 9493–9507.
series No. 10. Tulsa, OK: Society of Exploration Palmer, D. 1991. The resolution of narrow low-velocity
Geophysicists. zones with the generalized reciprocal method. Geophys.
Nuttli, O. & Whitmore, J.D. 1962. On the determination of Prospect., 39: 1031–1060.
the polarization angle of the S wave. Bull. Seism. Soc. Am., Palmer, I.D. & Taviolia, J.L. 1980. Attenuation by squirt
52: 95–107. flow in undersaturated gas sands. Geophysics, 45:
1780–1792.
Palmström, A. 1982a. Problems during construction of the
Vardø tunnel – a 2.6 km long submarine road tunnel. Proc.
O of 4th IAEG Congress, New Delhi, 4: IV.231–IV.244.
Palmström, A. 1982b. The volumetric joint count – a useful
O’Brien, P.N.S. & Lucas, A.L. 1971. Velocity dispersion of and simple measure of the degree of rock mass jointing.
seismic waves. Geophys. Prospect., 19: 1–26. Proc. of 4th IAEG Congress, New Delhi, 5: V221–V228.
O’Connel, R.J. & Budiansky, B. 1977. Viscoelastic proper- Palmström, A. 1996. Application of seismic refraction survey
ties of fluid-saturated cracked solids. J. Geophys. Res., 82: in assessment of jointing. Conf. on Recent Advances in
5719–5736. Tunnelling Technology, New Delhi, 15–22.
Oberti, G., Carabelli, E., Goffi, L. & Rossi, P.P. 1979. Study of Parra, J.O., Hackert, C.L. & Xu, P.-C. 2002, Character-
an orthotropic rock mass: experimental techniques, com- ization of fractured low Q zones at the Buena Vista Hills
parative analysis of results. Proc. of 4th Int. Congr. Rock Mech. reservoir, California. Geophysics, 67: 1061–1070.
(ISRM), Montreux: 2: 485–491. Rotterdam: Balkema. Patton, D.F. 1966. Multiple modes of shear failure in rock and
Oda, M., Yamabe, T. & Kamemura, K. 1986. A crack tensor related materials. PhD Thesis, University of Illinois, 282 p.
and its relation to wave velocity anisotropy in jointed rock Paul, R.J. 1993. Seismic detection of overpressuring and frac-
masses. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., turing: an example from the Qaidam Basin, People’s
23(6 ): 387–397. Republic of China. Geophysics, 58: 1532–1543.
Ogawa, Y. 1986. Geology. Ishikawa, No. 40, Ishikawa Soil Paulsson, B.N.P. & King, M.S. 1980. Between-hole acoustic
Incorporated Association. surveying and monitoring of a granitic rock mass. Technical
Ohkubo & Teresaki, 1971. Oyo Corporation. Technical Note note. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 17:
RP-479. 371–376.
Oliver, J. & Isacks, B. 1967. Deep earthquake zones, anom- Paulsson, B.N.P. 1984. A laboratory study of seismic velocities
alous structures in the upper mantle, and the lithosphere. and attentuation of host rocks for a nuclear waste repository.
J. Geophys. Res., 72: 4259–4275. Rock Mechanics in Productivity and Protection. Symp. on
Olofsson, B. & Kommedal, J. 2002. Facing the processing Rock Mechanics, 25. Evanston, Ill. 1063–1073. New York:
challenges of the Valhall 1998 3D 4C OBS data. Extended Society of Mining Engineers of AIME.
Abstract: Geophysics in a shared earth model. Symp., Paulsson, B.N.P., Cook, N.G.W. & McEvilly, T.V. 1985.
Norwegian Petroleum Society NPF, Kristiansand, 103–106. Elastic-wave velocities and attenuation in an underground
Olson, J. & Pollard, D.D. 1989. Inferring paleostresses from granitic repository for nuclear waste. Geophysics, 50(4):
natural fracture patterns: a new method. Geology, 17: 551–570.
345–348. Paulsson, B.N.P., Smith, M.E., Tucker, K.E. & Fairborn, J.W.
Olsson, O. (Ed.). 1992. Site characterization and validation – 1993. The use of cross-well seismology to characterize and
final report. Stripa Project 92–22. 364 pp. Stockholm: SKB. monitor a steamed oil reservoir. Comprehensive Rock
646 References

Engineering. Hudson, J.A. (ed.). Vol. 3. Rock Testing and Site Mechanics of Jointed and Faulted Rock, Vienna,
Characterization, 651–669. Oxford: Pergamon Press. Rossmanith, H.P. (ed.): 749–755. Rotterdam: Balkema.
Payne, S.S., Worthington, M.H., Odling N.E. & West, L.J. Potters, J.H.H.M., Groenendaal, H.J.J., Oates, S.J., Hake, J.H.
2005. A field investigation of the permeability dependence & Kalden, A.B. 1999. The 3D shear experiment over the
of seismic amplitudes. Seismic Q–observation, mechanisms, Natih Field in Oman. Reservoir geology, data acquisition
and interpretation. Workshop at 67th EAGE Conf., Madrid. and anisotropy analysis. Geophys. Prospect., 47: 637–662.
Extended Abstract: Q-01. Convenors: Hargreaves, Best, Prasad, M. 1988. Experimental and theoretical considerations
Lancaster. of velocity and attenuation interactions with physical para-
Peng, J.Y., Aki, K., Chouet, B., Johnson, P., Lee, W.H.K., meters in sands. PhD Thesis, Kiel University.
Marks, S., Newberry, J.T., Ryall, A.S., Stewart, S.W. & Prasad, M. & Manghnani, M.H. 1997. Effects of pore and
Tottingham, D.M. 1987. Temporal change in coda Q differential pressure on compressional wave velocity and
associated with the Round Valley, California earthquake of quality factor in Berea and Michigan sandstones. Geophysics,
November 23, 1984. J. Geophys, Res., 92(B5): 3507–3526. 62(4 ): 1163–1176.
Pepin, G.P., Gonzalez, M.E., Bloys, J.B., Loften, J.E., Prasad, M. 2002. Acoustic measurements in unconsolidated
Schmidt, J.H., Naquin, C.J. & Ellis, S.T. 2004. Effect of sands at low effective pressure and overpressure detection.
drilling fluid temperature of fracture gradient: field meas- Geophysics, 67(2): 405–412.
urements and model predictions. Paper 04-527, Gulf Prasad, M. 2003. Velocity-permeability relations within
Rocks ’04, 6th North American Rock Mechanics Symp. on hydraulic units. Geophysics, 68(1): 108–117.
NARMS, Houston: TX. Pratt, H.R., Swolfs, H.S., Brace, W.F., Black, A.D. &
Pérez M.A., Gibson, R.L. & Toksöz, M.N. 1999a. Detection Handin, J.W. 1977. Elastic and transport properties of
of fracture orientation using azimuthal variation of P-wave in situ jointed granite. Int. J. Rock Mech. Min. Sci. &
AVO response. Geophysics, 64(4 ): 1253–1265. Geomech. Abstr., 14: 35–45.
Pérez M.A., Grechka, V. & Michelena, R.J. 1999b. Fracture Press, F. 1966. Seismic velocities. Handbook of Physical
detection in a carbonate reservoir using a variety of seismic Constants. Clark, S.G. (Ed.). 195–218 Geological Society
methods. Geophysics, 64(4 ): 1266–1276. of America.
Pine, R.J. & Batchelor, A.S. 1984. Downward migration Price, D.G., Malone, A.W. & Knill, J.L. 1970. The applica-
of shearing in jointed rock during hydraulic injections. tion of seismic methods in the design of rock bolt systems.
Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 21(5): Proc. of 1st IAEG Congress, Paris. II: 740–751. Le Comité
249–263. Français de Géologie de l’Ingénieur.
Pininska, J. 1977. Correlation between mechanical and acoustic Pride, S.R. & Berryman, J.G. 2003. Linear dynamics of
properties of flysch sandstones. Int. Symp. on Geotechnics of double-porosity dual-permeability materials. II. Fluid
Structurally Complex Formations, Capri 1977. I: 387–394. transport equation. Phys. Rev. E 68, 036604, 10 pp.
Associazione Geotecnica Italiana. Pride, S.R., Harris, J.M., Johnson, D.L., Mateeva, A.,
Pistre, V., Plona, T., Sinha, B. & seventeen Schlumberger Nihel, K.T., Nowack, R.L., Rector, J.W., Spetzler, J.W.,
associates 2002. A New Modular Sonic Tool Provides Wu, R., Yamomoto, T., Berryman, J.G. & Fehler, M.
Complete Acoustic Formation Characterization. 43rd 2003. Permeability dependence of seismic amplitudes.
Annual Symp. on SPWLA, Osio, Japan. Extended Abstract. The Leading Edge, June 2006, 22(6 ): 518–525. Society of
Plona, T.J., Winkler, K., Sinha, B. & D’Angelo, R. 1998. Exploration Geophysicists.
Measurement of stress direction and mechanical damage Pride, S.R., Berryman, J.G. & Harris, J.M. 2004. Seismic
around stressed boreholes using dipole and microsonic attenuation due to wave induced flow. J. Geophys. Res., 109:
techniques. SPE/ISRM 47234 in Proc. of EUROCK’98, 1: 1–19. (for Section 13.3.2, end).
123–129. Pujol, J. & Smithson, S. 1991. Seismic wave attenuation in
Plona, T.J., Sinha, B., Kane, M.R., Shenoy, R., Bose, S., volcanic rocks from VSP experiments. Geophysics, 56(9):
Walsh, J., Endo, T., Ikegami, T. & Skelton, O. 2002. 1441–1455.
Mechanical-damage detection and anisotropy evaluation Purdy, G.M. 1982. The variability in seismic structure of
using dipole sonic dispersion analysis. 43rd Annual Symp. on layer 2 near the east pacific rise at 12°N. J. Geophys. Res.,
SPWLA, Osio, Japan. Extended Abstracts. 87(B10): 8403–8416.
Plona, T.J., Winkler, K., D’Angelo, R., Sinha, B., Purdy, G.M. & Detrick, R.S. 1986. Crustal structure of the
Papanastasiou, P. & Cook, J.M. 1997. Acoustic detection of Mid-Atlantic ridge at 23°N from seismic refraction studies.
stress-induced effects around a borehole. Int. J. Rock Mech. J. Geophys. Res., 91: 3739–3762.
Min. Sci. & Geomech. Abstr., Paper No. 290, 539–546. Purdy, G.M. 1987. New observations of the shallow seismic
Proc. of NYROCKS ’97. structure of young oceanic crust. J. Geophys. Res., 92(B9):
Podolski, R., Kosior, A. & Piotrowski, P.C. 1990. The results of 9351–9362.
applied seismic tomography to localization of the anom- Purdy, G.M., Christeson, G.L., Fryer, G.J. & Berge, P.A.
alies of velocity fields in coal seams. Int. Conf. on the 1991. Upper crustal structure of the East-Pacific Rise from
References 647

ocean bottom refraction experiments: interpretation and petrographic character of carbonate rocks. Geophysics, 49:
placement of a deep drill hole. Eos Trans. AGU 72: 262. 1622–1636.
(Abstract). Rasolofosaon, P.N.J., Rabbel, W., Siegesmund, S. &
Pyrak-Nolte, L.J., Myer, L.R., Cook, N.G.W. & Vollbrecht, A. 2000. Characterization of crack distribution:
Witherspoon, P.A. 1987a. Hydraulic and mechanical prop- fabric analysis versus ultrasonic inversion. Geophys. J. Int.,
erties of natural fractures in low permeability rock. Proc. of 141: 425–453.
6th Int. Cong. Rock Mech., Herget, G. & Vongpaisal, S. Rasolofosaon, P.N.J. & Zinszner, B.E. 2002. Comparison
(Eds.). V.1, 225–231, Rotterdam: A.A. Balkema. between permeability anisotropy and elasticity anisotropy of
Pyrak-Nolte, L.J., Myer, L.R., Cook, N.G.W. 1987b. Seismic reservoir rocks. Geophysics, 67(1): 230–240.
visibility of fractures. Proc. of 28th US Symp. on Rock Rathore, J.S., Fjaer, E., Holt, R.M. & Renlie, L. 1995.
Mechanics, Farmer, Daeman, Desai, Glass and Neumann. P- and S-wave anisotropy of a synthetic sandstone with
(Eds.): 47–56, Rotterdam: Balkema. controlled crack geometry. Geophys. Prospect., 43: 711–728.
Pyrak-Nolte, L.J., Myer, L.R. & Cook, N.G.W. 1990. Raymer, L.L., Hunt, E.R. & Gardner, J.S. 1980. An
Transmission of seismic waves across single natural frac- improved sonic transit time-to-porosity transform. SPWLA
tures. J. Geophys. Res., 95: 8617–8638. 21 Ann. Logging Symp., 1–12.
Pyrak-Nolte, L.J. & Nolte, D.D. 1992. Frequency dependence Rector III, J.W. 1995. Special Issue: crosswell methods: where
of fracture stiffness. Geophys. Res. Lett., 19(3): 325–328. are we, where are we going? Geophysics, 60(6 ): 629–630.
Pyrak-Nolte, L.J., Xu, J. & Haley, G.M. 1992. Elastic inter- Remy, J.-M., Bellanger, M. & Homand-Etienne, F. 1994.
face waves along a fracture: theory and experiment. Proc. Laboratory velocities and attenuation of P-waves in lime-
of 33rd US Symp. on Rock Mechanics, Tillerson & stones during freeze-thaw cycles. Geophysics, 59(2): 245–251.
Wawersik (Eds.): 999–1007. Rotterdam: Balkema. Renshaw, C.E. 1995. On the relationship between mechan-
Pyrak-Nolte, L.J. 1996. The seismic response of fractures and ical and hydraulic apertures in rough-walled fractures.
interrelations among fracture properties. 1995 Schlumberger J. Geophys. Res., 100(B12): 24,629–24,639.
Lecture Award Paper, Int. J. Rock Mech. Min. Sci. & Ribacchi, R. 1988. Rock mass deformability: in situ tests,
Geomech. Abstr., 33(8): 787–802. their interpretation and typical results in Italy. Int. Symp.
on Field Measurements in Geomechanics, 2. Kobe 1987. 1:
171–192. Rotterdam: Balkema.
Q
Roche, E. 2001. Seismic data acquisition – The new millen-
nium. Geophysics, 66: 54.
Quadros, E. 1982. Determinação das caracteristicas do fluxo
Roche, S.L., Davis, T.L. & Benson, R.D. 1997. 3C seismic
de água em fraturas de rochas. Dissert. De Mestrado,
study at Vacuum Field, New Mexico. Annual Int. Meeting
Department of Civil Engineering, Polytech. School,
Society of Exploration Geophysicists, Expanded Abstracts:
University of São Paulo, Brazil.
67: 886–889.
Quadros, E. 1995. Water percolation in rock formations in
Rodrigues, L.F., Oliveira, R. & Correia de Sousa, A. 1983.
Brazil. Special lecture. 8th ISRM Congress, Tokyo.
Cabril dam – control of the grouting effectiveness by
Workshop on Rock Foundation.
geophysical seismic tests. Proc. of 5th Int. Congr. Rock Mech.
Quadros, E. & Correa Filho, D.C. 1995. Grouting efficiency
(ISRM), Melbourne: 1: A1–A4. Rotterdam: Balkema.
using directional (3-D) hydraulic tests in Pirapora Dam,
Roecker, S.W., Tucker, B., King, J. & Hatzfeld, D. 1982.
Brazil. Proc. of 8th ISRM Congress, Tokyo. Fuji, T. (ed.). 2:
Estimates of Q in central Asia as a function of frequency
823–826. Rotterdam: Balkema.
and depth using the coda of locally recorded earthquakes.
Quan, Y. & Harris, J. 1997. Seismic attenuation and tomog-
Bull. Seism. Soc. Am., 72: 129–149.
raphy using the frequency shift method, Geophysics, 62(3):
Rohr, K.M.M. 1994. Increase of seismic velocities in upper
895–905.
oceanic crust and hydrothermal circulation in the Juan de
Queen, J.H. & Rizer, W.D. 1990. An integrated study of
Fuca plate. Geophys. Res. Lett., 21: 2163–2166.
seismic anisotropy and the natural fracture system at the
Rossi, G., Böhm, G., Gei, D. & Madrussani, G. 2005. Atten-
Conoco borehole test facility, Kay County, Oklahoma.
uation tomography: an application to gas-hydrate and free-gas
J. Geophys. Res., 95(B7 ): 11,255–11,273.
detection. Seismic Q–observation, mechanisms, and interpreta-
tion. Workshop at 67th EAGE Conference, Madrid. Extended
R Abstract: Q-03, Convenors: Hargreaves, Best and Lancaster.
Rowlands, H.J., Booth, D.C. & Chiu, J.-M. 1993. Shear-
Raestad, N. 2002. Barents Sea – geological possibilities, wave splitting from microearthquakes in the New Madrid
political complications. (in Norwegian) Geo-Bergverksnytt, seismic zone. Can. J. Explor. Geophys., 29: 352–362.
April 2002: 12–16. Trondheim. Rüger, A. 1996. Reflection coefficients and azimuthal AVO
Rafavich, F., Kendall, C.H.St.C. & Todd, T.P. 1984. analysis in anisotropic media. PhD Thesis, Colorado School
The relationship between acoustic properties and the of Mines.
648 References

Rummel, F., Alheid, H.J. & Frohn, C. 1978. Dilatancy and Savich, A.I., Koptev, V.I. & Zamakhaiev, A.M. 1974. In situ
fracture induced velocity changes in rock and their rela- ultrasonic investigation of failure of limestone. Proc. of 3rd
tion to frictional sliding. Pageophysik, 116: 743–764, ISRM Congress, Denver: Co IIA: 418–423. Washington:
Basel: Birkhäuser Verlag. DC, National Academy of Sciences.
Rutledge, J.T. & Winkler, H. 1987. Attenuation measure- Savich, A.I., Mikhailov, A.D., Koptev, V.I. & Iljin, M.M. 1983.
ments in basalt using vertical seismic profile data from the Geophysical studies of rock masses. Proc. of 5th Int. Congr.
eastern Norwegian Sea. Society of Exploration Geophysicsts, Rock Mech. (ISRM), Melbourne: 1: A19–A30. Rotterdam:
Expanded Abstracts: 711–713. Balkema.
Savich, A.I., Ilyin, M.M. & Yakoubov, V.A. 1987. Regularities
of alteration of elastic waves velocities in footing massif of
S Inguri arch dam during its operation. Proc. of 6th Int.
Congr. Rock Mech. (ISRM), Montreal, Herget, G. &
Saenger, E.H. & Shapiro, S.A. 2002. Effective velocities Vongpaisal (Eds.): 1: 515–519. Rotterdam: Balkema.
in fractured media: a numerical study using the rotated stag- Sayers, C.M. 1990. Stress-induced fluid flow anisotropy in
gered finite-difference grid. Geophys. Prospect., 50: 183–194. fractured rocks. Transport Media, 5: 287–297.
Saito, T. 1981. Variation of physical properties of igneous rock Sayers, C.M. 1998. Misalignment of the orientation of frac-
in weathering. Proc. of Int. symp. on Weak Rock, Tokyo, 1, tures and the principal axes for P and S waves in rocks con-
191–196. Rotterdam: Balkema. taining multiple non-orthogonal fracture sets. (Abstract)
Salisbury, M.H. & Christensen, N.I. 1978. The seismic Geophys. J. Int., 133: 459–466.
velocity of a traverse through the Bay of Islands ophiolite Sayers, C.M. & Dean, S. 2001. Azimuth-dependent AVO in
complex, Newfoundland, an exposure of oceanic crust and reservoirs containing non-orthogonal fracture sets. Geophys.
upper mantle. J. Geophys. Res., 83(B2): 805–817. Prospect., 49: 100–106.
Salisbury, M.H., Milkereit, B., Ascough, G., Adair, R., Sayers, C.M. 2002a. Stress-dependent elastic anisotropy of
Matthews, L., Schmitt, D.R., Mwenifumbo, J., Eaton, D.W. sandstones. Geophys. Prospect., 50: 85–95.
& Wu, J. 2000. Physical properties and seismic imaging of Sayers, C.M. 2002b. Fluid-dependent shear-wave splitting in
massive sulphides. Geophysics, 65(6 ): 1882–1889. fractured media. Geophys. Prospect., 50: 393–401.
Sammonds, P.R., Ayling, M.R., Meredith, P.G., Murrell, Schlumberger GeoQuest, UK Nirex Ltd. Report S/94/007,
S.A.F. & Jones, C. 1989. A laboratory investigation of by kind permission.
acoustic emission and elastic wave velocity changes during Schneider, B. 1967. New methods for the exploration of rock
rock failure under triaxial stresses. Rock at Great Depth, masses. (in French) Annales ITBTP, July/August, 235–236,
Proc. of Eurock, Maury, V. & Fourmaintraux, D. (Eds.): Paris, France.
233–240. Rotterdam: Balkema. Schoenberg, M. 1980. Elastic wave behaviour across linear
Sampaolo, A., Cello, G. & Moro, T. 1978. Preliminary slip interfaces. J. Acoust. Soc. Am., 68(5 ): 1516–1521.
results of a study relative to TBM-rock interaction. Proc. Schoenberg, M. & Sayers, C.M. 1995. Seismic anisotropy of
of 3rd IAEG Congress, Madrid. 10: 236–237. Madrid: fractured rock. Geophysics, 60: 204–211.
Imprime ADOSA. Schoenberg, M. 1998. Acoustic characterization of under-
Sams, M.S. 1995. Attenuation and anisotropy: the effect of ground fractures: 68th Annual Int. SEG Meeting, New
extra fine layering. Geophysics, 60: 1646–1655. Orleans, Expanded Abstract: 1624–1627.
Sams, M.S., Neep, J.P., Worthington, M.H. & King, M.S. Schrauf, T.W. 1984. Relationship between gas conductivity and
1997. The measurement of velocity dispersion and geometry of a natural fracture. MSc Thesis, Department of
frequency-dependent intrinsic attenuation in sedimentary Hydrology and Water Resources, University of Arizona,
rocks. Geophysics, 62(5): 1456–1464. 156 pp.
Sato, J., Itoh, J., Aydan, Ö. & Akagi, T. 1995. Prediction of Scott, J.H., Lee, F.T., Carroll, R.D. & Robinson, C.S.
time-dependent behaviour of a tunnel in squeezing rocks. 1968. The relationship of geophysical measurements to
FMGM ’95, 4th Int. Symp., Bergamo, Italy. 47–54. engineering and constructional parameters in the Straight
Sato, H. & Fehler, M.C. 1998. Seismic Wave Propagation and Creek Tunnel pilot bore, Colorado. Int. J. Rock Mech.
Scattering in the Heterogeneous Earth, 308 pp. New York: Min. Sci. & Geomech. Abstr., 5: 1–30.
Springer Verlag & AIP Press. Scott, T.E. & Nielsen, K.C. 1991. The effects of porosity on
Sattel, G., Frey, P. & Amberg, R. 1992. Prediction ahead of the the brittle-ductile transition in sandstones. J. Geophys. R.,
tunnel face by seismic methods. Pilot project in Centovalli 96: 405–414.
Tunnel, Locarno, Switzerland. First Break, 10: 19–25. Scott, Jr., T.E., Ma, Q. & Roegiers, J.-C. 1993. Acoustic veloc-
Savage, J.C. 1969. Comments on velocity and attenuation of ity changes during shear enhanced compaction of sandstone.
seismic waves in imperfectly elastic rock. (Gordon and Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 30(7 ):
Davis, 1968) J. Geophys. Res., 74: 726–728. 763–769. (34th US Symp. on Rock Mechanics, Madison WI.)
References 649

Scott, Jr., T.E., Ma, Q., Roegiers, C. & Reches, Z. 1994. Singh, S. & Herrmann, R.B. 1983. Regionalization of crustal
Dynamic stress mapping utilising ultrasonic tomography. coda Q in the continental United States. J. Geophys. Res., 88:
1st NARMS ’94. Austin TX. Rock Mechanics. Nelson & 527–538.
Laubach (Eds.). 427–434. Rotterdam: Balkema. Sisavath, S., Al-Yaarubi, A., Pain, C.C. & Zimmerman, R.W.
Sen, A. & Bandyopadhyay, M.R. 1990. Subsurface explor- 2003. A simple model for deviations from the cubic law
ation by seismic refraction for a pumped storage scheme in for a fracture undergoing dilation or closure. Pure Appl.
Northern India. Proc. of 6th IAEG Congress, Amsterdam. Geophys., 160: 1009–1022.
Price (ed.). 2: 1025–1031. Rotterdam: Balkema. Sjøgren, B., Øfsthus, A. & Sandberg, J. 1979. Seismic classifica-
Seraphim, J.L. & Pereira, J.P. 1983. Considerations of the geo- tion of rock mass qualities. Geophys. Prospect., 27: 409–442.
mechanics classification of Bieniawski. Proc. of Int. Symp. Sjøgren, B. 1984. Shallow Refraction Seismics, London,
on Eng. Geology and Underground Construction, LNEC, Chapman & Hall. 268 p.
Lisbon 1: II33–II42. Sjøgren, B. 2000. A brief study of the generalized reciprocal
Shakeel, A. 1995. The effect of oriented fractures on elastic method and of some limitations of the method. Geophys.
wave velocities, attenuation and fluid permeability of sand- Prospect., 48: 815–834.
stones. PhD Thesis, Imperial College, University of London. Skjærstein, A. & Fjær, E. 2000. Angular dependent attenuation
Shakeel, A. & King, M.S. 1998. Acoustic wave anisotropy in in a sandstone with parallel cracks. In: Advances in Aniso-
sandstones with systems of aligned cracks. In Core-Log tropy – Selected Theory, Modelling and Case Studies. Hood,
Integration, Harvey, P.K. & Lovell, M.A. (Eds.). 173–183. J.A. (Ed.). 205–215. Society of Exploration Geophysicists.
Geological Society of London, Special Publication No. 136. Slater, C., Crampin, S., Brodov, L.Y. & Kuznetsov, V.M.
Sharp, J.C. 1970. Fluid flow through fissured media. PhD 1993. Observations of anisotropic cusps in transversely
Thesis, 181 pp. University of London. isotropic clay. Can. J. Expl. Geophys., 29: 216–226.
Shaw, F., Worthington, M.H., Andersen, M.S. & Petersen, Slater, C. 1997. Estimation and modelling of anisotropy in
U.K. 2004. A study of seismic attenuation in basalt using vertical and walkaway seismic profiles at two North
VSP data from a Faroe Islands borehole. P 015, Extended Caucasus oil fields. PhD Thesis, Department of Geology
Abstract: EAGE 66th Conf. and Exhib, Paris. and Geophysics, University of Edinburgh.
Shaw, P. 1994. Age variations of oceanic crust Poisson’s ratio: Slimak, Š.A., Ljumovíc, G.R., Lokin, P.M. & Mašala, S.R.
inversion and a porosity evolution model. J. Geophys. Res., 1991. Determination of discontinuous rock mass properties
99(B2): 3057–3066. by seismo-acoustic methods. Proc. of 7th ISRM Congress,
Shea, V.R. & Hanson, D.R. 1988. Elastic wave velocity and Aachen. Wittke (Ed.). 1: 609–614. Rotterdam: Balkema.
attenuation as used to define phases of loading and failure Slotnick, M.M. 1936. On seismic computation with appli-
in coal. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., cations, I: Geophysics 1: 9–22.
25(6 ): 431–437. Smith, B., Guilbot, J. & Pirera, F. 2002. Time-lapse consistent
Shearer, P.M. & Orcutt, J.A. 1986. Compressional and shear depth conversion using filtered stacking velocities and hori-
wave anisotropy in the oceanic lithosphere – the Ngendei zontal wells at Ekofisk. Extended Abstract: geophysics in a
seismic refraction experiment. Geophys. J. Roy. Astr. Soc., shared earth model. Symp., Norwegian Petroleum Society
87: 967–1003. NPF, Kristiansand, 31–34.
Shearer, P.M. 1988. Cracked media, Poisson’s ratio and the Smith, R.L. & McGarrity, J.P. 2001. Cracking the fractures –
structure of the Upper Oceanic Crust. Geophys. J. Int., 92: seismic anisotropy in an offshore reservoir. The Leading
357–362. Edge, January 2001, 18–26.
Shen, B. & Stephansson, O. 1993. Numerical analysis of Mode Snow, D.T. 1968. Rock fracture spacings, openings, and
I and Mode II propagation of rock fractures. Int. J. Rock porosities. J. of Soil Mech. and Found. Div. Proc. of ASCE,
Mech. Min. Sci. & Geomech. Abst., 30(7 ): 861–867. SMI. 73–91.
Shen, B. & Barton, N. 1997. The disturbed zone around Solomon, S.C. & Toksöz, M.N. 1970. Lateral variation of
tunnels in jointed rock masses. Technical Note, Int. J. Rock attenuation of P and S waves beneath the United States.
Mech. Min. Sci. & Geomech. Abstr., 34(1): 117–125. Bull. Seism. Soc. Am., 60: 819–838.
Shen, B. Stephansson, O. & Rinne, M. 2002. Simulation of Somerton, W.H., Söylemezoglu, I.M. & Dudley, R.C. 1975.
borehole breakouts using FRACOD2D, In Oil and Gas Effect of stress on permeability of coal. Int. J. Rock Mech.
Science and Technology – Revue de l’IFP, special issue for Min. Sci. & Geomech. Abstr., 12: 129–145.
Int. Workshop of Geomechanics in Reservoir Simulation Spencer, J.W. 1979. Bulk and shear attenuation in Berea
57(5): 579–590. IFP. Rueil-Malmaison, France. sandstone: the effects of pore fluids. J. Geophy. Res., 84:
Simiyu, S.M. 2000. Geothermal reservoir characterization: 7521–7523.
application of microseismicity and seismic wave properties Spencer, J.W. 1981. Stress relaxation at low frequencies in
at Olkaria, Kenya rift. J. Geophys. Res., 105: fluid saturated rocks: attenuation and modulus dispersion.
13,779–13,795. J. Geophys. Res., 86: 1803–1812.
650 References

Spudich, P. & Orcutt, J. 1980a. A new look at the seismic Tanimoto, C. & Kishida, K. 1994. Seismic geotomography:
velocity structure of the oceanic crust. Rev. Geophys. Space amplitude versus velocity in consideration of joint aperture
Phys., 18: 627–645. and spacing. 1st NARMS 94. Austin, TX. Nelson & Laubach
Spudich, P. & Orcutt, J. 1980b. Petrology and porosity of (Eds.). 147–155. Rotterdam: Balkema.
an oceanic crustal site: results from wave form modelling of Terzaghi, K. 1943. Theoretical Soil Mechanics, New York:
seismic refraction data. J. Geophys. Res., 85(B3): 1409–1433. John Wiley & Sons.
Stacey, T.R. 1977. Seismic assessment of rock masses. Symp. Teufel, L.W. & Farrell, H.E. 1992. Interrelationship between
on Exploration for Rock Engineering. Johannesburg 1976, 2: in-situ stress, natural fractures and reservoir permeability
113–117. Rotterdam: Balkema. anisotropy – A case study of the Ekofisk Field, North
Stainsby, S.D. & Worthington, M.H. 1985. Q estimation Sea. Proc. of 1st Int. Conf. on Fractured and Jointed Rock
from vertical seismic profile data and anomalous variations Masses (ISRM Symp.), Vienna, Rossmanith, H.-P. (Ed.).
in the central North Sea. Geophysics, 50: 615–626. Thill, R.E., Tweeton, D.R., Jessop, J.A. & Roessler, K.S.
Stapledon, D.H. & Rissler, P. 1983. Site exploration and 1992. Developments in seismic tomography for minesite
evaluation. General report. Proc. of 5th Int. Congr. Rock investigation. Rock Mechanics, Proc. of 33rd U.S. Symp.
Mech. (ISRM), Melbourne: 3: G5–G25. Rotterdam: Rock Mech., Santa Fe, NM, Tillerson, J.R. & Wawersik,
Balkema. W.R. (Eds.): 1093–1102. Rotterdam: Balkema.
Stead, D., Szczepanik, Z. & Mackintosh, A.D. 1990. Borehole Thomsen, L. 1986. Weak elastic anisotropy. Geophysics, 51:
acoustic logging around underground potash mine open- 1954–1966.
ings. Proc. of 6th IAEG Congress, Amsterdam. Price (Ed.). Thomsen, L. 2002a. The future of 4C and 4D seismics: chal-
2: 1037–1043. Rotterdam: Balkema. lenges and needs. Extended Abstract: geophysics in a
Stenin, V.P., Kasimov, A.N. & Tikhonov, A.A. 2002. Cracked shared earth model. Symp., Norwegian Petroleum Society
layers characterization using far offset VSP. Proc. 10th Int. NPF, Kristiansand, 35–36.
Workshop on Seismic Anisotropy (IWSA), Tutzing. Thomsen, L. 2002b. Understanding seismic anisotropy
Strandenes, S. 1991. Rock physics analysis of the Brent in exploration and exploitation. 2002 Distinguished
Group reservoir in the Oseberg Field. Stanford Rockphysics & Instructor Short Course, Series No.5, Society of Exploration
Borehole Geophysics Project, Special Volume, 25 pp. Geophysicists and European Association of Geoscientists
Swift, S.A., Kent, G.M., Detrick, R.S., Collins, J.A. & and Engineers, (SEG/EAGE).
Stephen, R.A. 1998a. Oceanic basement structure sedi- Timur, A. 1968. Velocity of compressional waves in
ment thickness, and heat flow near Hole 504B. J. Geophys. porous media at permafrost temperatures. Geophysics, 33:
Res., 103(B7 ): 15,377–15,392. 584–595.
Swift, S.A., Lizarralde, D., Stephen, R.A. & Hoskins, H. Tingay, M., Müller, B., Reinecker, J., Heidbach, O.,
1998b. Velocity structure in upper ocean crust at Hole 504B Wenzel, F. & Fleckenstein, P. 2005. Understanding tec-
from vertical seismic profiles. J. Geophys. Res., 103(B7 ): tonic stress in the oil patch: The World Stress Map Project.
15,361–15,376. The Leading Edge, December 2005, 1276–1282.
Swolfs, 1977. Contract Report. TerraTek International, Salt Tjaaland, E., Kanestrøm, R., Folstad, P.G. & E. Fjær, 2001.
Lake City. Effects of overburden anisotropy and attenuation on seismic
Swolfs, H.S., Brechtel, C.E., Brace, W.F. & Pratt, H.R. 1981. inversion results for reservoir parameters. 9th IWSA (Int.
Field mechanical properties of a jointed sandstone. Mecha- Workshop on Seismic Anisotropy), Abstracts, Geophysics,
nical behaviour of crustal rocks, Geophysical Monographs, 66(4): 1294–1312.
American Geophysical Union, 24: 161–172. AGU Tod, S.R. 2002. The effects of cracks and porous flow on
seismic waves in crustal rock. PhD Thesis, Clare College,
University of Cambridge.
Tod, S.R., Hudson, J.A. & Liu, E. 2002a. Modelling
T frequency–dependent anisotropy due to fluid flow in bed
limited cracks. 72nd Annual Int. SEG Meeting, Salt Lake City.
Talebi, S. & Young, R.P. 1992. Microseismic monitoring Tod, S.R., Hudson, J.A. & Liu, E. 2002b. Pressure induced
in highly stressed granite: relation between shaft-wall anisotropy of interconnected cracks. Proc. of 10th Int.
cracking and in situ stress. Int. J. Rock Mech. Min. Workshop on Seismic Anisotropy (IWSA), Tutzing.
Sci. & Geomech. Abstr., 29(1): 25–34. Tod, S.R. & Liu, E. 2002. Frequency-dependent anisotropy
Talwani, M. 1964. A review of marine geophysics. Marine due to fluid flow in bed-limited cracks. Geophys. Res. Lett.,
Geol., 2: 29–80. 29: 15 (Paper No. 10.1029/2002GL015369).
Tanimoto, C. & Ikeda, K. 1983. Acoustic and mechanical prop- Todd, T. & Simmons, G. 1972. Effect of pore pressure on
erties of jointed rock. Proc. of 5th Int. Congr. Rock Mech. the velocity of compressional waves in low porosity rocks.
(ISRM), Melbourne: 1: A15–A18. Rotterdam: Balkema. J. Geophys. Res., 77: 3731–3743.
References 651

Toksöz, M.N., Cheng, C.H. & Timur, A. 1976. Velocities program leg 130, Ontong Java. J. Geophys. Res., 98(B5):
of seismic waves in porous rocks. Geophysics, 41: 7903–7920.
621–645. Ursin, B. & Toverud, T. 1985. Computation of zero-offset
Toksöz, M.N., Johnston, D.H. & Timur, A. 1979. seismic profiles including geometrical spreading and absorp-
Attenuation of seismic waves in dry and saturated rocks: tion. Geophys. Prospect., 33: 72–96.
laboratory measurements. Geophysics, 44: 681–690.
Toomey, A. & Bean, C.J. 2000. Numerical simulation of
seismic waves using a discrete particle scheme. Geophys. J. V
Int., 141: 595–604.
Torres-Verdin, C., Alpak, O.F., Wu, J., Gao, G.-Z., Hou, J., Van Der Baan, M. 2002. Constant Q and a fractal, stratified
Varela, O.J., Gambús-Ordaz, Chi, S. & Youmelin, E. earth. Pure Appl. Geophys., 159: 1707–1718.
2003. A multi-physics, integrated approach to formation Van der Kolk, C.M., Guset, W.S. & Potters, J.H.H.M. 2001.
evaluation using borehole geophysical measurements and The 3D shear experiment over the Naith field in Oman:
3D seismic data. Society Exploration Geophysicists, 73rd the effect of fracture-filling fluids on shear propagation.
Annual Int. Meeting, Dallas, TX. Extended Abstracts. Geophys. Prospect., 49: 179–197.
Toverud, T. & Ursin, B. 2005. Comparison of seismic attenu- Van Dok, R., Gaiser, J. & Probert, T. 2004. Time-lapse shear-
ation models using zero-offset vertical seismic profiling wave splitting analysis at Ekofisk Field. Paper G046, 66th
(VSP) data. Geophysics, 70(2): F17–F25. EAGE Ann. Conf. and Exhib., Paris, France.
Townend, J. & Zoback, M.D. 2000. How faulting keeps the Van Everdingen, D.A. 1995. Fracture characteristics of the
crust strong. Geology, 28: 399–402. sheeted dike complex, Troodos ophiolite, Cyprus:
Tsidzi, K.E.N. 1997. Propagation characteristics of ultrasonic Implications for permeability of oceanic crust. J. Geophys.
waves in foliated rocks. IAEG Bulletin, 56: 103–113. Res., 100(B10): 19,957–19,972.
Tsikalas, F., Gudlaugsson, S.T. & Faleide, J.I. 1998. The Vera, E.E., Mutter, J.C., Buhl, P., Orcutt, J.A., Harding, A.J.,
anatomy of a buried complex impact structure: the Mjølner Kappus, M.E., Detrick, R.S. & Brocher, T.M. 1990. The
structure, Barents Sea. J. Geophys. Res., 103(B12): structure of 0- to 0.2-m.y.-old oceanic crust at 9°N on the
30,469–30,483. East Pacific Rise from expanded spread profiles. J. Geophys.
Tsujiura, M. 1978. Spectral analysis of the coda waves from Res., 95: 15,529–15,556.
local earthquakes. Bull. Earthquake Res. Inst., 53: 1–48. Vidstrand, P. 2003. Äspö Hard Rock Laboratory. Update
Univ. Tokyo. of the hydrogeological model 2002. Int. Progress Report
Tsvankin, I. & Lynn, H.B. 1999. Special section on IPR-03-35.
azimuthal dependence of P–wave seismic signatures – Vlastos, S., Liu, E., Main, I.G. & Narteau, C. 2005. Seismic
introduction. Geophysics, 64: 1139–1142. wave propagation in fractured media: scattering attenuation
Turk, N. & Dearman, W.R. 1986. A suggested approach to as an indicator of fracture network evolution. Seismic
rock characterization in terms of seismic velocities. 27th US Q–observation, mechanisms, and interpretation. Workshop
Symposium on Rock Mechanics, Tuscaloosa, AL. Hartman at 67th EAGE Conf., Madrid. Extended Abstract: Q-07,
(Ed.): 168–175. Littleton, CO: Society for Mining, Convenors: Hargreaves, Best and Lancaster.
Metallurgy and Exploration (SME). Vlastos, S., Narteau, C., Liu, E. & Main, I. 2002. Numerical
Tutuncu, A.N. 1998a. Nonlinear viscoelastic behaviour of study of scattering attenuation in fractured media: the
sedimentary rocks, Part I: effect of frequency and strain effects of scale length on multiple scattering attenuation.
amplitude. Geophysics, 63: 184–194. Abstract 73rd Annual Int. SEG Meeting, Dallas, TX.
Tutuncu, A.N. 1998b. Nonlinear viscoelastic behaviour of Vlastos, S., Narteau, C., Liu, E. & Main, I. 2003a. Numerical
sedimentary rocks, Part II: hysteresis effects and influence of study of scattering attenuation in fractured media: frequency
type of fluid on elastic moduli. Geophysics, 63: 195–203. dependence and effects of characteristic length scales.
Tutuncu, A.N., Podio, A.L., Gregory, A.R. & Sharma, M.M. Abstract EAGE 65th Conf. and Exhibition, Stavanger.
1998. Nonlinear viscoelastic behaviour of sedimentary Vlastos, S., Schoenberg, M., Maillot, B., Main, I., Liu, E. &
rocks, Part I: effect of frequency and strain amplitude. Li, X.-Y. 2003b. Fluid flow simulation in a fractured
Geophysics, 63(1): 184–194. medium: effects of pore pressure changes on seismic
waves. Abstract EAGE 65th Conf. and Exhibition,
Stavanger.
Vogelaar, B.B.S.A. & Smeulders, D.M.J. 2005. Low
U frequency attenuation in a saturated rock. Seismic
Q–observation, mechanisms, and interpretation. Workshop
Urmos, J. & Williams, R.H. 1993. In situ velocities in at 67th EAGE Conference, Madrid. Extended Abstract:
pelagic carbonates: new insights from ocean drilling Q-05, Convenors: Hargreaves, Best and Lancaster.
652 References

Volti, T. & Crampin, S. 2003. A four-year study of shear- White, R.S. 1984. Atlantic oceanic crust: seismic structure of
wave splitting in Iceland: 2. Temporal changes before a slow spreading ridge. In: Ophiolites and Oceanic
earthquakes and volcanic eruptions. In: New Insights into Lithosphere. Glass, Lippard & Shelton (Eds.). 34–44.
Structural Interpretation and Modelling. Nieuwland, D.A. Geological Society of London.
(Ed.). Geological Society of London Special Publication, White, R.S. & Whitmarsh, R.B. 1984. An investigation of
212. 135–139. seismic anisotropy due to cracks in the upper oceanic crust
at 45°N, Mid-Atlantic ridge. Geophys. J. Roy. Astr. Soc., 79:
439–467.
W Whiteley, R.J. 1990. Keynote lecture: advances in engineering
seismics. Proc. of 6th IAEG Congress, Amsterdam. Price
Walsh, J.B. 1966. Seismic wave attenuation in rock due to (Ed.). 2: 813–825. Rotterdam: Balkema.
friction. J. Geophys. Res., 71: 2591–2599. Whitmarsh, R.B. 1978. Seismic refraction studies of the upper
Walsh, J.B. 1995. Seismic attenuation in partially saturated igneous crust in the North Atlantic and porosity estimates
rock. J. Geophys. Res., 100(B6 ): 15,407–15,424. for Layer 2. Earth Plan. Sci. Lett., 37: 451–464.
Wang, Z. & Nur, A. 1990. Wave velocities in Hydrocarbon- Wilcock, W.S.D., Solomon, S.C., Purdy, G.M. & Toomey,
saturated rocks: experimental results. Geophysics, 55: D.R. 1992. The seismic attenuation structure of a fast-
723–733. spreading Mid-Ocean ridge. Science, 258: 1470–1474.
Wang, Z. & Nur, A. 1992. Seismic and acoustic velocities in Wilcock, W.S.D., Solomon, S., Purdy, G.M. & Toomey, D.R.
reservoir rocks. Geophysics reprint series. No. 10. Wang, & 1995. Seismic attenuation structure of the East Pacific Rise
Nur (Eds.). Society of Exploration Geophysicists. near 9°30N. J. Geophys. Res., 100(B12): 24,147–24,165.
Watanabe, T. & Sassa, K. 1996. Seismic attenuation tomog- Wilkens, R., Simmons, G. & Caruso, L. 1984. The ratio
raphy and its application to rock mass evaluation. Int. J. Vp/Vs as a discriminant of composition for siliceous lime-
Rock Mech. Min. Sci. & Geomech. Abstr., 33(5): 467–477. stones. Geophysics, 49: 1850–1860.
Wei, K. & Liu, D. 1990. The zoning system for assessment Wilkens, R., Schultz, D. & Carlson, R. 1988. Relationship
of weathering states og granites. Proc. of 6th IAEG of resistivity, velocity and porosity for basalts from down-
Congress, Amsterdam. Price (Ed.). 1: 657–663. hole well-logging measurements in hole 418A. Proc. of
Rotterdam: Balkema. Ocean Drilling Project, Scientific Results, Salisbury, M.H.,
Weiss, R.F., Lonsdal, P., Lupton, J.E., Banbridge, A.E. & Scott, J.H., et al. (Eds.). 102: 69–75.
Craig, H. 1997. Hydrothermal plumes on the Galapagos Wilkens, R.H., Fryer, G.J. & Karsten, J. 1991. Evolution of
rift. Nature, 267: 600–603. porosity and seismic structure of Upper Oceanic Crust:
Wenhua, Z. 1991. Treatment of fracture zone in Gezhouba Importance of aspect ratios. J. Geophys. Res., 96(B11):
project. Int. Congress on Large Dams, 17. Vienna 1991. 3 17,981–17,995.
(Q66): 1667–1679. Paris: ICOLD. Williams, M. & Jenner, E. 2002. Interpreting seismic data in the
Wepfer, W.A. & Christensen, N.I. 1987. Characterizing presence of azimuthal anisotropy; or azimuthal anisotropy in
microcracks via a velocity-pressure relation. Eos, Trans. the presence of the seismic interpretation. The Leading Edge,
AGU, 68, 1503. August 2002, also CSEG Recorder, June 2003, 35–39.
Wepfer, W.W. & Christensen, N.I. 1990. Compressional wave Willis, M.E., Rethford, G.L. and Bielanski, E. 1986.
attenuation in oceanic basalts. J. Geophys. Res., 95: Azimuthal anisotropy: occurrence and effect on shear-
17,431–17,439. wave data quality. Presented at the 56th Annual Int.
Wepfer, W.W. & Christensen, N.I. 1991. A seismic velocity- Meeting. Society of Exploration Geophysicists.
confining pressure relation, with applications. Technical Willis, M.E. & Rao, R. 2005. Seismic attenuation losses
note. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 28(5): induced by discrete vertically aligned fractures. Seismic
451–456. Q–Observation, Mechanisms, and Interpretation. Workshop at
Westerdahl, H. 1998. Private communication. 67th EAGE Conference, Madrid. Extended Abstract: Q-13,
Westerdahl, H. & Cosma, C. 1998. Private communication. Convenors: Hargreaves, Best and Lancaster.
Westerman, A.R., Reeves, M.J. & Attwell, P.B. 1982. Rock Willson, S.M. & Fredrich, J.T. 2005. Geomechanics consid-
surface roughness and discontinuity closure at depth. erations for through- and near-salt well design. Paper
Developments in Geotechnical Engineering, 32: 63–67, SPE 95621, SPE Annual Tech. Conf. and Exhibition,
Elsevier. Dallas: TX.
Westman, E.C., Haramy, K.Y. & Rock, A.D. 1996. Seismic Willson, S.M. 2006. Feeling the heat, can’t stand the pressure?
tomography for longwall stress analysis. 2nd NARMS ’96. Risks to borehole integrity when drilling in ultra-HPHT
Montréal, Québec. Aubertin, Hassani & Mitri (Eds.). 1: environments. First Break, 24 May 2006, 37–41, EAGE.
397–403. Rotterdam: Balkema. Winkler, K. & Nur, A. 1979. Pore fluids and seismic attenu-
White, J.E. 1983. Underground Sound, New York: Elsevier. ation in rocks. Geophys. Res. Lett., 6: 1–4.
References 653

Winkler, K.A., Nur, A. & Gladwin, M. 1979. Friction and Worthington, M.H., King, M.S. & Marsden, J.R. 2001. Deter-
seismic attenuation in rocks. Nature, 277: 528–531. mining the damping factor of sedimentary rocks required
Winkler, K.W. & Nur, A. 1982. Seismic attenuation: effects for seismically designed structures. Int. J. Rock Mech. Min.
of pore fluids and frictional sliding. Geophysics, 47: 1–15. Sci. & Geomech. Abstr., 39: 801–806.
Winkler, K.W. 1983a. Contact stiffness in granular porous Worthington, M.H. & Lubbe, R. 2003. The scaling
materials. Comparison between theory and experiments. of fracture compliance. Geological Society, Special
J. Geophys. Res., 10: 1073–1076. Publication.
Winkler, K.W. 1983b. Frequency dependent ultrasonic Wulff, A.-M., Hashida, T., Watanabe, K. & Takahashi, H.
properties of high-porosity sandstones. J. Geophys. Res., 1999. Attenuating behaviour of tuffaceous sandstone and
88: 9493–9499. granite during microfracturing. Geophys. J. Int., 139:
Winkler, K.W. 1985. Dispersion analysis of velocity and 395–409.
attenuation in Berea sandstone. J. Geophys. Res., 90: Wulff, A.-M. & Mjaaland, S. 2002. Seismic monitoring of
6793–6800. fluid fronts: an experimental study. Geophysics, 67(1):
Winkler, K.W. & Murphy III, W.F. 1995. Acoustic velocity 221–229.
and attenuation in porous rocks. In: Rock Physics and Phase Wyllie, M.R.J., Gregory, A.R. & Gardner, L.W. 1956. Elastic
Relations: A Handbook of Physical Constants, American wave velocities in heterogeneous and porous media.
Geophysical Union, 1995. Geophysics, 21: 41–70.
Winkler, K.W., Sinha, B.K. & Plona, T.J. 1998. Effects of
borehole stress concentrations on dipole anisotropy meas-
urements. Geophysics, 63(1): 11–17. X
Winterstein, D.F., De, G.S. & Meadows, M.A. 2001. Twelve
years of vertical birefringence in nine-component VSP Xia, G. 2006. Fracture detection from seismic P-wave
data. Geophysics, 66: 582–597. azimuthal AVO analysis: application to Valhall LOFS data.
Witherspoon, P.A., Amick, C.H. & Gale, J.E. 1977. Stress- In situ rock stress. Proc. of Int. Conf. Trondheim, Lu, Li,
flow behaviour of a fault zone with fluid injection and Kjorholt & Dahle (Eds.). London: Taylor & Francis Group.
withdrawl. Mineral Engineering Report. No. 77-1,
University of California, Berkeley.
Witherspoon, P.A., Amick, C.H. & Gale, J.E. 1979. Stress-flow
behaviour of a fault zone with fluid injection and withdrawl. Y
Mineral Engineering Report No. 77-1, University of
California, Berkeley. Yale, D.P. & Jamieson, Jr., W.H. 1994. Static and dynamic
Witherspoon, P.A., Wang, J.S.Y., Iwai, K. & Gale, J.E. 1979. rock mechanical properties in the Hugoton and Panoma
Validity of cubic law for fluid flow in a deformable fracture. fields. Kansas Soc. Petr. Engr., Paper 27939.
Lawrence Berkeley Lab, LBL-9557, SAC-23, 28 pp. 1980 Yamamoto, S., Kimura, Y., Takahashi, K. & Miyajima, K.
Water Resour. Res., 16: 1016–1024. 1995. Geological investigation and evaluation of foundation
Wolfe, C.J., Purdy, G.M., Toomey, D.R. & Solomon, S.C. bedrock for long-span bridges in Japan. Rock Foundations.
1995. Microearthquake characteristics and crustal velocity Proc. of Int. Workshop on Rock Foundations, Tokyo, Yoshinaka,
structure at 29°N on the Mid-Atlantic ridge: the architec- R. & Kikuchi, K. (Eds.): 177–184. Rotterdam: Balkema.
ture of a slow spreading segment. J. Geophys. Res., Yamamoto, T. & Kuru, M. 1997. Imaging the hydro-
100(B12): 24,449–24,472. geological structure of the limestone aquifers of South
Won, G.W. & Raper, R.W.J. 1997. Downhole geophysical Florida using the super cross-well tomography: a pilot
investigations for a proposed deep highway cutting adjacent experiment using wells BF-1 and BF-2: University of
to a rail tunnel at Murrurundi, NSW, Australia. Modern Miami, Geo-Acoustics Laboratory Tech. Report No. 1038.
Geophysics in Engineering Geology. McCann, D.M. et al. Yamamoto, T. 2003. Imaging permeability structure within
(Eds.). 283–291. Geological Society Engineering Geology. the highly permeable carbonate earth: inverse theory and
Special Publication 12. experiment. Geophysics, 68(4 ): 1189–1201.
Wong, J., Hurley, P. & West, G.-F. 1983. Crosshole seismol- Yan, J. 2003. Improved rock physical models for integration
ogy and seismic imaging in crystalline rocks. Geophys. Res. of core, log and seismic data. PhD Thesis, University of
Lett., 10(8): 686–689. Edinburgh.
Wong, J. 2000. Crosshole seismic imaging for sulphide ore- Yielding, G., Walsh, J. & Watterson, J. 1992. The prediction
body delineation near Sudbury, Ontario, Canada. Geophysics, of small-scale faulting in reservoirs. First Break, 10:
65(6 ): 1900–1907. 449–460.
Worthington, M.H. & Hudson, J.A. 2000. Fault properties Young, R.P., Hill, J.J., Bryan, I.R. & Middleton, R. 1985.
from seismic, Q. Geophys. J. Int., 143: 937–944. Seismic spectroscopy in fracture characterization. Q. J.
654 References

Eng Geol., 18: 459–479. Northern Ireland: The Zeng, Y., Su, F. & Aki, K. 1991. Scattering wave energy
Geological Society. propagation in a medium with randomly distributed
Young, R.P. & Collins, D.S. 1999. Monitoring an experi- isotropic scatterers, 1. Theory. J. Geophys. Res., 96: 607–619.
mental tunnel seal in granite using acoustic emission and Zimmerman, R.W. & King, M.S. 1985. Propagation of
ultrasonic velocity. Rock Mechanics for Industry, Proc. of acoustic waves through cracked rock. 26th US Symp. on
37th U.S. Symp. Rock Mech., Vail, CO, Amadei, B., Kranz, Rock Mechanics, Rapid City SD. Ashworth (Ed.).
R.L., Scott, G.A. & Smeallie, P.H. (Eds.): 869–876. 739–745. Rotterdam: Balkema.
Rotterdam: Balkema. Zinszner, B., Meynier, P., Cabrera, J. & Volant, P. 2002.
Young, R.P. & Baker, C. 2001. Microseismic investigation of Vitesse des ondes ultrasonores, soniques et sismiques dans
rock fracture and its application in rock and petroleum les argilites du tunnel de Tournemire. Effet de l’anisotropie
engineering. Proc. of 4th Int. Workshop on Application of et de la fracturation naturelle. Oil and Gas Science and
Geophysics to Rock Engineering, Beijing, China. Int. Soc. Technology – Rev. IFP, 57(4): 341–353.
Rock Mech. News J., 7(1): 19–27. Ziolkowski, A. 1999. Chasing 1-meter scale with multi-well,
Young, R.P. & Collins, D.S. 2001. Seismic studies of rock multi-level, multi-component receivers. Offshore, 59(10):
fracture at the underground research laboratory, Canada. 74–76.
Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 38: Ziolkowski, A., Hanssen, P., Gatliff, R., Jakubowicz, H.,
787–799. Dobson, A., Hampson, G., Li, X.-Y. & Liu, E. 2003. Use of
Yu, G., Vozoff, K. & Durney, F.W. 1993. The influence of low frequencies for sub-basalt imaging. Geophys. Prospect.,
confining pressure and water saturation on dynamic elas- 51: 169–182.
tic properties of some Permian coals. Geophysics, 58(1): Zoback, M.D. & Hickman, S.H. 1982. In situ study of the
30–38. physical mechanisms controlling induced seismicity at
Yuan, J. 2001. Analysis of four-component seafloor seismic Monticello Reservoir, South Carolina. J. Geophys. Res., 87:
data for seismic anisotropy. PhD Thesis, University of 6959–6974.
Edinburgh. Zoback, M.D. & Townend, J. 2001. Implications of hydrostatic
pore pressures and high crustal strength for the deformation
of intraplate lithosphere. Tectonophysics, 336: 19–30.
Z Zoback, M.D., Barton, C.A., Brudy, M., Castillo, D.A.,
Finkleiner, T., Grollimund, B.R., Moos, D.B., Pesk, P.,
Zatsepin, S.V. & Crampin, S. 1997. Modelling the compli- Ward, C.D. & Wiprut, D.J. 2003. Determination of stress
ance of crustal rock. I Response of shear-wave splitting to orientation and magnitude in deep wells. Int. J. Rock Mech.
differential stress. Geophys. J. Int., 129: 477–494. Min. Sci. & Geomech. Abstr., 40: 1049–1076.
Index

ABC method 15 -quartz mixture (sand) 254


ABEM method 15 velocity of 12, 159
acid-leaching (model) 470, 480, 481 Alaska 228
acoustic Alaskan lithosphere 243
barrier 30 East-Central Alaskan crust 244
emission (see AE) albedo (see earthquakes)
closure of joints 3, 77–80, 166, 369 Alberta 386
dipole logging 301, 302, 312–319 aleurolite 212
impedence (see impedence) Alford rotation 314
log correction, mud-infiltration 317 for fast-shear tool azimuth 314
monopole logging 301, 318 for minimizing cross-receiver energies 314
sensors 59 aluminium 198, 424
acquisition geometry 7 alluvial
acute incident angle clay 26
w.r.t. bedding, foliation in shale 354, 355 stream-beds 375
adiabatic (dynamic) 183 alluvium 12, 142, 374
advance rate (TBM) 150, 158 Alpe Gera dam 77
AE acoustic emission 53, 85, 87, 128, 454 alum shale
activity 222 crushed 53, 139
arrays 130–131 fractured 162
inaudible to- 435 alteration 316
interpreted velocities 129–131 chemical near-wellbore- 316
subsidence causes 454 due to discontinuum formation 302–307
temporal changes 130 mechanical wellbore- 316
tomography 131 zone, time-dependent interactions 319
Aegean 220 alterered
Afar triangle, Ethiopia (thinnest crust) 245 shale 313
Afghanistan 228 shale compressional arrival 313
African Plate 230 un- undamaged formation 313
age, basement age un- virgin formation compressional
at mid-Atlantic ridge, and VP 262, 264, 265 arrival 313
at mid-Atlantic ridge, and porosity 263 zone 303, 305, 307, 308, 310, 311, 313, 316
combined mid-Atlantic, East Pacific data 287–294 zone development with time 313
effect on Qseis 228 alteration
geological 369–371 deep- zone 312
agglomerate 25 drilling mud induced- 312
Ahmedabad, India 219 zone phenomena 311
air discussion re influence on logging 311
-dry samples 22, 23 alternating hard and soft rocks 301, 304, 311
filled 30, 119 Amazonian Basin 231
guns 254, 261, 438 ambient and heated tests 514, 515, 527
gun arrays (P-wave sources) 394, 396 Ambiesta, Italy, limestones in situ 105
gun travel time data 258 American units 296–298, 320
large- guns, low frequency sub-basalt imaging 403 Amoco 331
656 Index

amphibolite 13, 32, 38, 70 detection 3, 382


compositionally homogeneous 249 double-effect 35
-gneiss contrasts (reflectors) 249 effects 35, 382
amplitude/magnitude elastic 360
ratio A/Ao 85, 86 fault zone effect 429
ratio versus joint frequency, roughness, clay-filling 85, 86 following grouting 170
of waves (with, without fractures) 199, 204 frequency-dependent- 472, 473
versus frequency: jointed samples 199 hexagonal 436
versus frequency: bedded coal 201 horizontal permeability- 382
versus time, S1, S2 reversal 455 hydraulic 360
amplitude of roughness (a) compared to length (L) maxima for slate, mica quartz schist, phyllite 248
508, 510 natural-, layering 387
ANDRA near-surface model 225
Andes 72, 231 permeability- 357–364, 445, 446
andesite 26, 28, 29, 70, 72, 73 Poisson’s ratio-, due to fracturing at geothermal field
Andean Belt 231 394–396
anelasticity of matrix 182 P-waves, bedding fractures 499, 500
angle of incidence P-waves in failing sample 540
effect on reflection coefficients 388 P-wave velocity- correlation with gas productivity 384
in AVO 388 productivity direction connection 450
in one- and two-set models 393 rotation of velocity anisotropy axis (blasting effect) 119
Angolan offshore exploratory well 312 rotation of velocity anisotropy axes-
angular discrepancies re assumed correlated phenomena saturation reduction effect 480
polarization, anisotropy, non-parallel ␴Hmax shear wave-, supposed small range 413, 415
direction 384, 385, 389, 390, 403, 404, shear wave-, actual larger range 415
406, 414, 429, 431, 432, 475, 518 theory, basic 373, 374
anhydrite, VP and density 326 three classes of-, elastic tensors 374
anisotropic 9 three-dimensional 38, 39, 355–364
attenuation 354 velocity anisotropy 119, 357–364
brittle fracturing 36 VP–for 26 rock types at 1 GPa confinement 248
cuspidal phases 450, 451 with respect to assumed ␴Hmax 382, 383, 385,
domains 445, 446 390, 404, 406
dilation 36 anisotropy caused by
frequency 384 azimuthal shear modulus 404
jointing 9 cleavage 38, 248
pattern of wave attenuation 356 conjugate joint, fracture sets 392, 393
permeability 355, 359, 360 crack content 445, 446
permeability, Äspö spiral tunnel 555, 556 fabric 38, 248, 263
pore-pressure propagation due to fractures 464 faults, faulting 47, 48, 394
reflectivity 384 fluid-filled cracks, Mid-Atlantic Ridge 436, 437
response 304 foliation 38, 248
seismic velocities 3 fractures, aligned 358–360, 382, 436, 437, 439, 445,
stiffness 3 446, 525
stresses 35, 36, 298–300, 305, 358 fractures, aligned, poro-elastic models 461–481
stress influence on joint shearing 299 fractures, aligned, poro-elastic physical models
structure (EDA microcracks) 432, 428–435 480, 481
velocity 354 joints and bedding 361–364, 372, 373
anisotropy 40, 41, 42, 351–364, 394 heterogeneity 373
attenuation- 118, 382 horizontal (bedding-plane) fractures 499, 500
attenuation in presence of- 351–364 horizontal stress anisotropy 382, 439
azimuthal dependence of- 248 lower-frequency shear waves 459
azimuthal- from dipole sonic logging 391 micro-cracks 35
azimuthal P-wave- 405 micro-cracks in failing sample, P-wave- 540
azimuthal velocity- 35, 248, 382 oriented RQDo and oriented Jr/Ja 411
Index 657

over-burden 440 kinematic- 441


permeability 382 lenticular-, fluid-bearing lense 414, 525–527
principal stress difference 129 physical (E) 162, 172, 173, 278, 363,
rock joints in situ 35, 40, 41, 42 364, 392
stress difference 35 reduced with depth 276
subsiding overburden, joint stretch/stress/strain 453 scaling for aperture and spacing , micro- and
up-stepping, down-stepping 2nd joint set 523, 524, macro- 441
551, 552 sealed by post-kinematic cements 441
anisotropically -squared 173
loaded 306–311 stress-closure measurements of- 527–530
fractured 463, 465 stress-closure modelling of- 392, 531–534
anomalies unstressed aperture, equation, discussion 511, 532
break-out-, not ␴Hmax related 541 Appalachian Basin 370
fracture zones, sub-ocean 272, 273 aquaduct 79, 80
high density- 254–257 Arabian Plate 230
Poisson’s ratio 270 Archangelsk Region, Russia 439
temperature, conductive fractures 541, 542 area/volume ratio of pore space 33
velocity- 14, 256, 257 areal well shoot 44
velocity-age reversal 274, 275 Argentina 163, 244
anorthosites Arphy dam site, France 159
gabbroic- 242 artificially jointed samples 40, 41
massive 147 aspect ratio 207
meta- 147 apertures e and E (see aperture) 467
anoxic, sub ocean floor 281 dominant 208
anticlinal structures 381 elliptical-crack closure stresses 529
jointing, fracturing caused by- 381 influence in modelling saturation effects 326–329
anticlinal trap 391 low- cracks 200, 201, 270, 276, 363, 364, 467
Anza, California 222 low- cracks sealed 270, 276
APE model, Crampin, critique of 415 low- cracks closed 287
aperture (fracture/joint) 456 low- fractures 391
aspect ratio aspect of- 392 means micro-cracks are stiffer than joints 412
closure unless strong rock, rough joints 392 of cracks and fissures 266
conducting (e) – normal stress, block test 514 of fractures, joints 363
conducting (e) – normal stress – temperature, block test of microcracks 226
515–517 pore shape spectra 326
conducting (e), modelled 309, 310 asperity (ubiquitous term)
cubed 161–164, 173 crushing 501, 502
e E 470, 480 role in dilation 519, 520
e E experimental data 457, 488–490 -weld, aluminium 424
e E demonstration re grouting 467, 468, 487, 490 welded-, Lucite, supposition 418
e E demonstration re tunnel model 469 Äspö, Sweden 119, 131–133
e E UDEC-BB modelling 460, 519–521 access spiral tunnel, permeability 555, 556
e E relevance to poro-elastic modelling 392, 456, 457, Asthenosphere 227, 241
467, 474 asymmetry of S- and P-wave reflections 438
e E relevance to production rates 453, 456, 487 Atlantic
e E relevance to shearing 474 Coastal Plain, USA 231
effect of JRC on-, reservoir BB simulation 533 (North-) fracture zones 273, 274
equation linking e and E by JRC 364, 487, 489 margin 244, 246, 251, 255, 256
extremely small-, reservoir BB simulations 531, 532 Ocean 244, 246, 251, 255, 256
‘frozen’-, by mineral infill 278 Shield, South America 231
hydraulic, conducting (e) 162, 172, 173, 278, 363, 364, sub-ocean Mid-Atlantic ridge 261–273
392 atomic
increase linked to dilation 524, 526 inter- spacing 188
indicated by mineralization (ophiolite) 278 weights 23
658 Index

attenuated 8 sensitivity to saturating fluid 386


attenuating separation (intrinsic and scattering attenuation) 213, 214
effect of cracks 184, 185 source of- 5
effect of cracks nucleating 463 spectrum 118
effect of freeze-thaw cycles 184, 185 structure, sub-ocean ridge area 282
joints support more flow 425 ‘total’- 192
layer 66 valid mechanism of- 189
attenuation (see 1/Qseis, 1/Q at Qseis, Q ) viscosity-based- 209
across continents 226–231 attenuation due to, (note 1/Q and Q considered)
amplitude 66 alteration-mineralogy 280
-anisotropy 119, 352, 481 axial strain 195–197
azimuthal- anisotropy 405 bulk modulus 356
coefficient 233, 353 clay, clay content 345, 346
data from sub-ocean sites and lab samples 280 clay-rich, clay-free sandstones 349
difference 118 compliant mineral content 347
expressions, after Thomsen 356 confining pressure, (effective) 191–193, 201
extreme, faulted rock, tunnel collapse 558 crack density 440
fluid-flow based- 186 depth (near-surface) 202, 203
for distinguishing lithologies 369 depth (1 km) 222, 223, 234
four components of- 193–194 depth (few km) 222, 232, 233, 237
high and low- regions (plate tectonics) 227 depth (many km) 229, 230
intrinsic 213, 214, 223, 280, 387 depth (extreme, plate-tectonic setting ) 179, 227
intrinsic-, incorrectly assessed 379 differential pressure, stress level 186, 187
intrinsic link to joint processes 490 distance from epicentre 211, 216, 220, 221
intrinsic squirt flow (see squirt flow) 236 double-porosity limestones 348–350
influence of Ko on- 299, 300 dry rock, perfectly dry 343
in presence of anisotropy 351–357 effective stress change 386
lateral variations 228 E-modulus of specimen (supposition) 191–193, 196, 200,
low, Sugar Loaf, Rio de Janeiro 558 201, 342
mineral/fabric related 246–248 fault zone 235, 427, 428
mechanisms 17, 182, 183, 195, 196, 345 formation stiffness variations 386
magnitudes 299 fractures, fracture frequency 224, 280, 286
maximum 353 fracture compliance 408
measurements in petroleum reservoirs 232–238 fracture nucleation 463
mechanisms discussion, earthquakes 219, 220 frequency of freeze/thaw 185
mechanisms, Coulomb friction, squirt, scattering 282 frequency of coda Q (earthquakes) 211, 214, 217, 223,
melt-squirt- 242 228
near-surface mechanisms of- 224 frequency (ubiquitous topic) 187, 189, 190, 194, 200,
of velocity components 363 207, 208, 238, 343–345
P-wave- 382–396, 459, 481 frequency of measurement (VSP, X-hole, sonic, core) 208,
peak- versus frequency 387 209
peak (frequency  viscocity rule) 207 friction 186, 187–189, 195–197, 206, 219, 224
peak (aspect ratio range) 207 friction, calculated 197
peaks at low and full saturation 190 geology/lithology (ubiquitous topic) 206, (222)
problems 49, 51, 57 geometrical spreading losses 386
profiles 387 heterogeneities 191
profile (post-blast) 118 high porosity and high permeability 349
profile (pre-blast) 118 inelasticity of matrix 387
reduced at low frequency, sub-basalt 403 increased fracturing 280
rotation of attenuation anisotropy axis 119 inter-bedded units when imaging sub-basalt 403
scattering 182, 195–197, 213, 214, 223, 280 intrinsic links to macro-processes 487
scattering mechanisms 202 lateral location in crust 229
sensitivity to structure 386 magnitude, seismic 220
sensitivity to effective stress change 386 normal stress ( joints) 199
Index 659

over-pressure 350, 351 Austin, Texas 444


porosity 186 Austria 76, 95
pore pressure change 193 Australia 25, 67, 200, 201
reflections 386 Aveyron, France 39
rock mass quality Q (see J- components) 200–203, 210, AVO 369, 388–392, 407
220, 221, 224–226, 350 anomalous most positive, most negative gradients
rough basalt surface 403 391, 392
saturating fluid 386 attributes, fracture parameters relation 389
saturation with water 186, 189, 190 gradients, azimuthal anisotropy effects 389
saturation with gas/brine 341 gradient, maximum 390
shear mobilization, friction 507 most positive near-offset- gradient 392
shear modulus 356 response with additional fracture sets 393
shear strain 189 AVOA 369, 388–392
squirt flow (see squirt) 17, 182, 183, 187, 190, 202, P-wave amplitude 388
206, 346, 349, 355, 356, 457, 461, 465–476 azimuth 314
stress-aligned micro-cracks 394 binning 384, 452
strain amplitude 187–189 fast- changing across fault 315
strain amplitude and confinement 188 full- 3D P-wave survey recommendation 384
structure 386 narrow- data 373
temporal seismic events (before/after) 215, 216, 218 sectoring 374
viscous losses in compliant pores 387 stacking 452
attenuation, intrinsic 182, 195–197 azimuthal
absorption 210 anisotropy (see anisotropy)
mechanisms 202, 206 anisotropy attributable to structure controlling
attenuation, inverse of Q (see 1/Qseis, 1/Q at Qseis and Q, all fluid flow 382
categories of data) anisotropy in subsiding overburden 453, 454
1/Q (1/QP ) – frequency data comparison using Biot 209 bins of data 384, 437
1/Q (1/QP ) – axial stress-strain, loaded to failure 195, 196 deviation of S-wave polarization due to O/R
1/Q (1/QP ) – frequency, calculated for fractures /joints of 548–552
different stiffness 200 deviation of S-wave polarization due to unequal conjugate
1/Q (1/QP ) – frequency, nucleating, coalescing fracture net- joints 549–552
works 463 isotropy (TIV) 451
1000/QP – frequency dependence(interbedded sand-, lime-, sensitivity, with single set 393
silt- and mudstones) 207 separation, of reflection coefficients 393
1000/QE – extensional strain amplitude 187 variation of amplitudes 447
1000/QE and 1000/Qp – confining pressure 191 variation of travel times 447
Q1 or 1/Q or 1000/Q or Qseis1 7, 9, 45, 79, 80, 179, velocity anisotropy 3, 40–43, 314, 317, 373, 382
Ch10: 181–239, Ch13: 344, 345, 349, 363 velocity anisotropy signalling aligned fracturing 382
attenuation recorded during earthquakes velocity anisotropy signalling anisotropic stress 382
albedo (Bo) 214 azimuthally
Q1 actually (Qc1) intrinsic (earthquakes) 213, 214 differentiated seismic attributes 384
Q1 actually (Qc1) scattering (earthquakes) 213, 214 isotropic source and receiver arrays 384
QS1 intrinsic component 213, 214, depth-dependence limited seismic attributes 384
213 axis of ridge (see mid-ocean)
Qsc1 (or ScQS1) scattering component 214 axial
total 214 modulus (␺) 5, 13
attenuation recorded in fractured reservoirs strain 195–197, 300, 339
depth – Q1 (and Q), in fault zone 235 strain amplitude 340
depth – Q1 – VP (interbedded sediments) 234, (380) stress (see VP) 300, 339
depth – Q1 – VP –K (fractured shale) 237 valley 437
(Q in anticlinal chalk) 235 valley floor (mid-ocean ridge) 436, 437
attenuation recorded in sub-ocean ridge areas (see Qseis listings) Azerbaijan reservoir, deeply buried sandstones 312
QS – depth (hole 504B, Costa Rica ridge, 0 to 1600 m, also Äspö, Sweden 119, 131–133
Qintrinsic) 280 access spiral tunnel, permeability 555, 556
660 Index

Bad Creek borehole, permeability 555 basic friction angle (see friction)
Balder formation 15 basin (see also plate tectonics) 179
band width (ubiquitous term) forearc 227
very high- with down-hole instruments 378 marginal 227
Bakhtar calculator curves 526 pull-apart 228
Ballotini glass spheres 547, 548, 551, 552 Basin and Range province, Western USA 227, 231
Bandis joint strength, scale-effect components 522 Bay of Islands, Newfoundland 262, 264
Barents Sea 404 bathymetric contours 281, 437
Barton-Bandis constitutive model for rock joints, fractures beach sand 254, 350, 351
401, 424, 428, 435, 456, 484, 486, 490, 492, beam footprints 67
493, 537 Beaumont Tunnel, near Channel Tunnel, S.England
application if clay-fillings, caution 546, 547 94, 319
conversion of strength criterion to Mohr-Coulomb beautiful
parameters 539 strikingly- (continuous sub-ocean seismic profile) 284
coupled modelling 506 Beaver County, Utah 42
coupled stress, deformation, dilation, permeability 523, bed, bedding
524, 526 intra- joints 208, 236
high-stress version, confined JCS 544 intra- joint apertures 208
modelling of conducting apertures under stress joints 145
512, 541 -parallel 352–357
modelling of conducting apertures, closure and shear -perpendicular 352–357
531–533 plane opening 319, 320
modelling of saturation on shear stiffness 511, 512 plane slip, subsidence 399, 400, 453, 454
Barton-Choubey shear strength criterion 27, 428, 537 thickness
application if clay-fillings, caution 546, 547 before-and-after-fracturing VP VS Q1 and permeability
conversion to Mohr-Coulomb parameters 539 362–364
peak shear stiffness estimate 428 Bekkelaget, Norway 160
basalt 12, 20, 21, 25, 28, 29 bell-shaped curve 190
artificial surface in- 80 BEM FRACOD models 306, 307
Atlantic margin- 255 bentonite
breccia 277 seals 129–130
Columbia Plateau- 269 Berea (see sandstone)
Columbia River, dilated joint 520 Berger, Norway 160
columnar- 124–127, 366, 500 Berkeley, California 422
columnar/flow entablature- , Hanford 124–127 BGS Anisotropy Project, Edinburgh 405, 409
compliance and dynamic stiffness, columnar 500, 503, BHA bottom hole assembly 312
504 BHC borehole compensated logging 379
Eastern North Sea- 269 BHTV 385, 386
flows 277 biaxial
flow-top weathering 163 loading of in situ blocks 165–167, 513–517
flow-top permeability 163 loading of tension fracture model 105, 524
intact, distinction from jointed 269, 276 loading of 1 m3 blocks 509, 545, 553
North Atlantic, Faroe Islands 203 loading-path, theoretical, corrected for dilation 553
permeable, jointed 270 biaxially-loaded direct shear box
pillow 277 needing correct stress transformation, dilation added 554
sub- imaging 403 bi-compressional arrival 313, 317
-to garnet transformation 242 phantom arrival 313
unfractured 278 billiard-table planarity, non-producers 549
vesicular 22, 23, 265 Biot 183
basalt, mid-ocean 243, 255, 264, 267–270, 273, flow related to shear modulus 356
277, 280, 285 fluid flow (boundary shear) 183, 349
basement 8, 221 fluid-past-frame attenuation 356
crystalline 220, 222 -Gassmann fluid substitution, mud-filtration 317
sub-ocean 244 generalized theory of poro-elasticity 320
Index 661

poroelasticity equations 209 break-out, log-spiral 302–307


theoretical prediction of 209 break-out, extensional 303
biotite gneiss break-out, slotted 303
highly weathered 11 break-out (stress) 250, 252, 299, 301, 302–311
jointed 11 causing dispersive effects 314
sound 11 causing frequency dependent effects 314
weathered 11 -compensated sonic logging BHC 47
birefringence (see shear wave splitting) 445, 446 creep 297, 299
bit size variation at KTB 251 damaged zone 301–311
Blake Spur, North Atlantic fracture zone profile 273 deep-, conducting fractures study 536, 541, 542
blast/blasting 30 deviated 305
gasses 133 disturbed zone 301–311
induced fracturing 63, 64 effects 271
normal- 127, 128 elongation 472
post- 63 ellipticity 301
pre- 63 enlargement due to overbreak 299
smooth- 127, 128 extensometers, multiple position, MPBX 301
blastability 17 fluid velocity 206
block, blocks (rock) heaters 513
block tests, lab, large-scale 509, 553 heater test 134–136
inter-block friction coefficient 6 images 471
failure 167 instrumented- 137, 303, 305
falls 304 large-diameter 133, 307, 309
heated- test, CSM /Terratek 513–517 large-diameter analogue 319
length, size sheared 493 loading test 136–138
modelled- 306, 308–310 logging 249, 252
natural- size 422 logging of disturbed zone 312–320
of Berea sandstone 137, 315 over-break (jointing) 250, 252, 299, 307–310
of diverse sandstones 359–364 plastic flow 299, 303
relative block-size 6 pre-drilled- 302
samples 27 radial stress gradients 314
scaling JRC, JCS to –size 401 research- for seismic studies (see Hole) 207, 208, 299
size effect on stress-strain behaviour, models 523, 524 shallow test- 203
size, increasing 308 shear or extension failures 299
sizes L1 L2 L3 524 size reduction 299
size, relative (see RQD/Jn) 225 stress anisotropy 137, 314
size scaling of Ks 421, 422, 523 stress concentrations 136–138, 314, 386
tension fracture- models 524, 551 stress concentration reasons for VSP sonic mismatch 386
test, in situ 165–167, 513 squeezing 297, 299
test UDEC-BB model 518, 519 televiewer logs, logged 277, 384, 385, 541, 542
Blow-Me-Down massif, Bay of Islands 264 to seismic spread 14
Bluebell Altamont Field, Uinta Basin, Utah velocity 70, 73
471–475, 525 wall condition 295, 299, 302–311
Blyvoor gold mine, South Africa 131 Borrego Valley, southern California 260
bolting 621 boulders 12
bolt spacing 621 boundary condition
BOSK effective medium model 457 one-dimensional strain 392, 401
fracture porosity 457 roller- 401
fracture density 458 Boussinesque elastic foundation 162
gas or brine distinction 457, 458 BP Claire Field 404–406
boring 30 BP 4D-4C plans 394
borehole, boreholes 3, 9, 11 BP Devine Test Site, USA 234
ballooning 321 BP Valhall reservoir 453
break-out (see break-out) Brazil 173, 174
662 Index

Brazilian contrasts, Sugar Loaf, tunnel collapse 558 Cajon Pass scientific borehole, San Bernadino, California
Brazilian dam foundation testing 505 432, 541, 554
Brazilian Shield 231 mobilized friction calculations 544, 554
break-out permeabilities 555
analysis 383, 386, 432 calcarenite 100
anomalies 541 calcite 509
borehole- (‘dog-earing’) 302–307 -filled joints 79, 165, 167
log-spiral shear-failure surfaces 130, 302–307 calcium carbonate cement 206
orientation logs 389 caldera structure, Kenya Rift 394–396
perpendicular to- 390 Caledonian quartzites 11
phenomena, theoretical 303 California 21, 22, 23, 210, 212, 386, 429, 431,
stress-controlled-, vertical stress 319 432, 541
structurally-controlled- 319 Southern 228
V-shaped notches (tunnels, shafts) 127, 128, 131, 303 Californian earthquakes 226
wedge-shaped fall-out 304 calliper log
breccia, brecciated rock 161 four-arm 382, 472
Brenntangen, Norway 160 long axis parallel ␴h min 382, 383
bridges, long span 19 measurements 250, 252, 304
bright spots 458 noise 320
brine Cambourne School of Mines, Cornwall 169
-filled cracks 326, 327, 329, 426, 457, 458, 479, 480 Cambrian sandstones 11
saturation effects 326–329 Cambro-Silurian limestones 11
brittle 37 Canada 22, 57, 61, 87, 117, 127–131, 217, 325, 331,
failure 369–391, 505, 506
-ductile transition 36, 37, 257, 311, 312, 537, 539 cap rock 372
shear fracturing 312 capillary effects 353
broad-band dispersion capillary pressure (see patchy saturation)
curves 316 carbonate, carbonates
of dipole flexural waves 316 beds 387, 388
of Stonely wave modes 316 domal 443
broken rock 118 Palaezoic 430
buckling mode of deformation 310, 311 rocks 196
Buena Vista Hills reservoir, California 236, 387, 388 Carboniferous 447
Bulgaria 19 limestone 499, 500–503
bulk modulus (K) (see also deformation) 5, 13, 71, 104, sandstone 11
109–111, 529 siltstone 11
of infill material, relation to ZN /ZT compliance ratio casing
425, 426 collapse 399
pressure-sensitive-, using excess compliance 397 collapse due to overburden stretch 399–401
reduction due to joint frequency 522, 523 damage 454
buoyancy effect 297 hundreds of- collapses 400
burial setting of- 297
present deep- 300 waiting for – installation 320
present shallow- 300 cataclastic
Bussesundet, Norway 145 fault seals 404
BWIP, Hanford, USA 124–127 flow of sandstones 312
Byerlee (frictional-strength) ‘law’ 536, 542 Catalina Island, southern California 260
Caucasan oil reservoirs 439, 450–452
C-waves 369, 388, 389 cavern, caverns 16, 43, 53, 55, 76, 123, 224
repeated 3C x 3C, 9C survey 451 reinforcement and support needs 93
cable CDR compensated dual resistivity tool 313
bottom drag 11 celluloid Qseis value 181
car, Rio 558 cement, cemented
Cabril dam site, Portugal 170 partially-, vuggy 404
Index 663

cementation 26 chlorite 509


high degrees of 370 fillings, outwashed 163
episode 334, 336 Christensen and Mooney
Central California earthquakes 213, 214 global review of velocity-depth structure 244–251
Central Europe 107 circular openings 303, 305, 307, 308–311
Cerro Prieto Geothermal Field (CPGF) 435, 436 circulation loss
chalcopyrite 23 civil engineering 3, 24, 161, 390
chalk 7, 12, 53, 54, 76 projects 14, 45, 324, 484
and cheese 369 Claire Field, UK offshore, Shetland Islands
artificial surfaces 80 attenuation anisotropy investigations 404–406
Austin-, Texas 444 multi-azimuth walk( float)-away VSP 404
base-, Qseis 235 classic
chalk marl, jointed 16, 77, 79 equations 4
Cretaceous, jointed 203, 204 relationships 369–372
domal 443 stress transformation equations 552
density-VP 325 stress transformation equations, modified 553
dual-porosity, low Q 350 classic knee-shaped curve
Eocene 123 for VP – pressure lab data 31, 32, 80, 81, 85, 87, 168,
foundation 29 192, 193, (198, analogue), 201, (250)
fractured 297, 391 for VP-depth trends, in situ, on-land and sub-ocean
intact 30 91, 93, 157, 271, 272, 286, 290, 166, 167,
jointed 94 241, 245, 248, 250, 251, 252–254, 258,
jointed, porous Ekofisk- 339–402, 411, 420, 428, 443, 260–262, 264, 266, 267, 274–277, 280, 281,
451, 452, 522, 549 284–286, 289
low quality 30 classification (of rockmass conditions) 144
Lower- 77 schemes (various sources) 19, 21, 26, 28, 69, 71–73, 75,
marl 17, 77, 79, 94, 163, 319 75, 106, 142–144, 148
melange 30 system (Q) (see also Q listings) 3, 13, 75, 92, 93, 108,
ooze transition 206 134, 144, 146, 150, 152, 164, 615–626
overlying- 234 classification of fractured reservoirs 333–335, 440, 441
pore collapse matrix porosity versus fracture porosity, 441
porous 79 Types I, II, III, IV 440, 441
porosity - VP 325 clastics
remoulded 30 Palaezoic 430
slickensided joints (see jointed, Ekofisk above) clay 7, 8, 11, 12
stress-conducting aperture, CSFT 529, 530 and attenuation 345, 345
thin- beds 206 and gravel 24
top-, Qseis 235 bearing discontinuities 271, 346, 493
Upper-Cretaceous 447 bearing, well-jointed, QH2O – K estimation 557
VP – density 326 coatings 148
VP – VS 328 dark- layers, effect on permeability 359
water-weakening, Ekofisk 400 filled discontinuities, break-down of Lugeon-1/Q
Chamoli earthquake, Garhwal Himalaya, India 218 relation 555
channel, fast velocity 15 filled discontinuities, displacements to peak 546
Channel Tunnel, between England and France 17, 94, 319 filled discontinuities, stiffnesses 427, 428, 505
channelling 160 fillings 169, 170, 543
characterization (see classification) 4, 11, 144, 615 fillings (artificial) 81
Charlie Gibbs, North Atlantic fracture zone profile 273 interlayers 101, 102, 212, 234
Cheddar Gorge 118 filled joints 505
chert, Mesozoic 153 low permeability- 268
Chevron 386 Maicop 450
Chile 72 Oxford- 49, 51
China 74, 76, 94, 110, 210, 212, 213, 245 platelet alignment 451
Chinnor Tunnel, Oxfordshire 16, 76, 163 -rich materials, flow 302
664 Index

clay (contd ) columns (see basalt)


/sand sequences 7 artificially fractured 81–86
-seal, -smear along faults 295 collapse, Brazilian tunnel 558
sealing of joints, absence of 210, 366 comb-gauge for roughness measurement 510
silty-sandy- 238 Comité National Français 159
-smear 308, 428 commercial prospectivity from neural network 384
soft/sensitive 53 commercially
swelling- important azimuthal anisotropy 446
silty- 9 non-viable fractures, low JRC 364
viscous interaction between- particles 194, 345 viable fractures, high JRC 364
clay content 17, 331–334 commonly held viewpoint re Hmax and flow
differentiation 331–335 direction 383
- porosity 331–334 compaction, compacting
- permeability 332–334 band (borehole failure mode) 303
- sonic contra seismic detection 332 drive (Ekofisk) 444
stratigraphy-guided 332–335 link to time shift 398, 402
volumetric- content, relation to QP 346 mechanical-, of porous sandstones 311, 312
claystone, Triassic, Lower Jurassic 427 modelling with UDEC-BB, Ekofisk 400, 401, 420
Climax mine, USA 135 monitoring in 4D, Ekofisk 398, 399
closed-form elasto-plastic analysis 305 one-dimensional strain- modelling 399–401
closure (see normal-closure, stress-closure) reservoir, conjugate jointing 420, 485
of microcracks 31 rubble-ization (Ekofisk) 443
CMP number, lateral 381 shale, South China sea 301
CO2 – flooding Companie Général de Géophysique CGG 396
effect of increased pore pressure 451 comparison of (ubiquitous term)
effect on reduced velocity 451 core and sonic-log, VP – pressure (depth) 348
Coachella Valley, southern California 260 competent (little weathered) 11
coal compliance, compliances (of cracks, joints, fractures, faults)
bedded- 201 5, 6, 98
bituminous (Pittsburgh, Pocahontas, Greenwich) 168 additions, to account for joint sets 393
cleated 168 assumptions 204
in-seam measurements in- 200 changes 309
mine, opencast 118 cyclical- in inter-beds 300
outbursts 60 dynamic 6
pillars 60, 61 discussion 282
-seams dynamic joint- 184
-seams, shallow 67 elastic- tensor 416, 417
Tower colliery 200, 201 estimated from block test (ZN) 515
triaxial 103, 104 excess-, due to fractures 417, 461
uniaxial 103, 104 fault zone 427
cobbles 12 fluid-type linkage 425, 426
coefficient of friction (see friction, Jr, Ja) 469 fracture, joint normal- (ZN) 6, 200, 204, 391, 417–419,
cockroaches 373 463, 465, 492
cohesion (see Mohr-Coulomb) fracture, joint shear- (ZT) 6, 391, 417–419, 463, 465,
dissipation of- 304 492
cold water (see water, water-flood) high- voids, propped by fluid pressure 412
Cole-Cole coupling from dielectrics 194 Hudson compliance diagonal (scale-effect model) 503, 504,
Colombia River Basalts, Washington State, 515, 526
USA 520 increase with scale 503, 504
Colorado, USA 139–142, 165, 167 in situ- in fault zone 427, 428
Colorado School of Mines CSM, experimental mine laboratory and field data compared 503
513–515 link to porosity, supposition 484–486
Columbia Plateau, Washington State 269 link to shear-wave splitting, theory 416–417
columnar joints in basalt (see basalt) model 236
Index 665

model for fault 236 ␴1–␴3, Mohr circle diameter 534, 537, 539, 540,
non-linear, stress dependent 464, 465 544, 545
normal-, sub-vertical structure or shear waves 408, 417 confining pressure 22, 31
off-diagonal component ZNV 506 artificially elevated 265, 269, 281
ratio ZN /ZT 417–426, 463, 465 extreme 1 GPa 246
ratio ZN /ZT of clay-filled discontinuities 506 versus porosity, extreme pressures 312
ratio ZN /ZT relation to aspect ratio, Liu model 426 conglomerates
ratio ZN /ZT (normalized) relation to fluid-type, Liu tunnel EDZ in- 121
model 426, 502 weathered 148
ratio ZN /ZT variation with stress, Pyrak-Nolte 426 conjugate
saw-cut surfaces, limestones, 5–60 MPa load 501–503 fractures 105, 414, 420, 524, 550
saw-cut surfaces, limestones, inverted dynamic stiffnesses jointing, as alternative to EDA 435
501, 502 joint shearing 414, 443, 444, 450, 550
shear and normal- relation 407 pair, sub-parallel to dominant set 390
shear stress application, effect on- 506, 507 set of joints 130
stiffness comparison 421–425, 487, 492, 493 steeply-dipping sets 442
theories for resolving fluid type 425, 426 conjugate joint shearing
units inversion to compare with stiffness 424, 427 causing contrary rotation of O and R 414, 527,
versus fracture scale (logging method), Worthington et al. 550, 551
503 connected
with off-vertical fractures or incident waves 419, 420 dis- small-scale fissures 359
ZN field measurements 499, 500 connectivity 356
ZN and ZT excess compliance matrix locations 417 joint- needs at geothermal site 436
ZN and ZT equality, or not, discussion 417–419, Conoco Borehole Test Facility 377, 384, 385, 445, 446,
463, 465 447
compliant consolidated rock layers 4
and non- pores 356 consolidation 9, 26
bonds 204 consolidation effects 4
cracks 207 constant thickness slices 374
minerals, attenuation 346, 347 continental divide 140
compressibility (see fluid) converted waves
compressional P to S termed C 394, 438
and shear wave seismic 7 mode- with TIV and TIH 394
bi- arrivals 313, 317 contact area (joint, fracture walls) 79, 80
bodywaves 4, 5 ratio Ao/A1 80
slowness 316 continuum
waves (ubiquitous term) 355, 357, 362 modelling 129, 135
compression wave becomes a discontinuum 302–307
amplitude 80–86 Conrad discontinuity 241
reflection 7 continent
refraction 7 cross- scale 179
compressive strength (see UCS) continental
uniaxial 13, 20, 24, 26, 27 crust 241–252, 255
of joint walls (see joint wall compressive strength, JCS) crystalline- crust 263, 263
compressive strength – porosity (shale) 26 borderland 260
concrete blocks 27 intruded- crust 255
roughness profiles 509 margins 254–261
conducting joints, fractures, (see permeability, joints, fractures) shelf 242
some limitations of conventional assumptions 518 sub-crust, velocity-depth structures 241, 242,
conductivity (see permeability) 245–251
confidence limits 171 continuity
confined loss of- 304
artificially 268 controversial points
strength JCS at reservoir depth 534 regarding effective stress 263, 265–267, 269
666 Index

controversial points (contd ) sizes 130


regarding sonic-log velocity oscillations in inter-beds sizes, maximum plausible 196, 197
300, 301 tensor technique 40, 41, 42
regarding stiffness and compliance magnitudes 282 thick- 270
regarding stress transformation 252–254 thin- 270
contraction when shearing 303, 304, 444 vertical- 445
conversion interface 438, 439 crack density 127, 130
converted P-S waves (see P-S) ambiguity 411, 413, 468, 469
cooling-joints (see joints) critical, description 444
copper critique of geophysics definition 411, 468, 469
Qseis 181 critique of limited range 412–415, 434, 442
core (see drill-core) 70 decay model, Tod 462
analysis 249, 499 Ekofisk chalk example 442
Inner-, Earth 241 extreme (Lucite plates) 427
length 13, 75 sum of two sets 442
oriented- 384, 385 theoretical 195, 196
Outer- 241 unrealistic values (Lucite laminates) 421
sampled 339 crater 249
samples 128, 207–209, 334, 525 creep 297, 299
side-wall 441 dislocation climb, salt, HPHT 321
sub- samples 357 undefined mechanism, salt, HPHT 321
ultrasonic VP and Qseis on- 380 Cretaceous
core logging 9, 12, 76, 115, 623, App.A 615–624 and Eocene, SW Texas 370
field sheet 622 chalk 203
statistics 624 limestone 11
value if intersecting structure 407, (617) sandstone-shale units 371
Core (film) 182 critical shearing crust concept 542–552
coring, slant-hole 377 conducting/non-conducting fractures (Zoback and others)
Cornwall, England 40 527, 534, 541–544
borehole data 543 crack density description 444
geothermal project 413, 453, 536 mobilization of friction in crust 542–544
cost pre-peak or post-peak friction mobilization 545–548
of construction 141 shear strength of crust, non-linear 536–540
of support and grouting (relative) 146 state line, Barton 537, 539
Costa Rica ridge area, eastern equatorial Pacific criticality
277–279 caused by microcracks, critique of Crampin viewpoint
Coulomb friction 282 412, 413
coupled behaviour 165, 166, 487–492 -normality comparison, re tunnel stability, leakage 413
shear-dilation-flow (see CSFT) 474, 488–492, 507 cross-hole logging (see also tomography) 45–46, 49–52, 57,
Coyne et Bellier 110 160, 207–209
crack, cracks, cracking, cracks per meter 13, 71, 72 cross-hole
closure 35 model simulation 204
closure pressure 22 reflection imaging 67
closure under stress, Walsh 364, 528, 529 seismic 50–52, 66, 106, 117, 119, 120, 125, 380
closure-opening hysteresis 31, 32 seismic, for joint compliance ZN 499, 500
compliance (see compliance) 6 seismic, theoretical aspects 477
critical- density critique 413 seismic tomography (see tomography) 374–378, 504
dipping- 445 cross-well
dry- 43 reflection tomography 66
extension- 130 seismic tomography and permeability 377, 378
percentage of- 273 attenuation 234
relaxation mechanism 387 seismic 203
saturated 43 seismic (high frequency) 377, 387
sealed 270 seismic, shallow 377
Index 667

seismic Q 234 Dacite 28, 29


high-resolution seismic, Qseis and VP 236, 237 dam 3, 339
transmission 66 abutment 170, 173, 174
cross-correlation of parameters 21 arch- 49–51, 88–90, 99, 102
cross-discipline parallel 310, 333 foundations 50, 51, 57, 99, 174, 505
cross-discipline region (wellbore alteration zone) 319 projects 49–51
cross-over of P-wave and S-wave attenuation 343 sites 27, 56, 77, 101, 102–109, 114, 115, 119, 163, 170
Cromer Knoll Group 447 site canyons 50, 51, 56, 99
crushing due to blasting 61 damage
crushed zones 145 zone (see excavation damage zone EDZ)
at borehole scale 306, 308 zone confirmed by bi-compressional arrivals 317
CRUST 5.1 crustal velocity-depth model 252 crack-related- (freeze-thaw cycles) 185
crust, crustal (see Earth’s crust) 228 damaged rock 130, 303–310, 313, 317
age 228, 262–265, 287–294 damping (see attenuation)
anisotropy 246–248 Darcy (see permeability)
large- faults 223 data acquisition, test equipment
lower 249 field tests (cross-hole, tomography) 62, 125
middle 249 field test (3D permeability) 174, 292
old 270 lab tests (rock physics) 59, 192, 197, 361
phenomena 179 deep, deeply
velocity histograms 247 layer velocities 140, 141, 156
velocity histograms, depths of 5, 10, 15, 20, 25 km 247 weathered 4
thickness: mean, thinnest, thickest 245, 247 wells in hard crystalline rocks 538, 541–543
thinner, warmer- 223 well behaviour 179
upper- velocity-age relation, sub-ocean 287–289 Deep Sea Drilling Project DSDP 261, 276, 278
young 270 Deere 76
crystalline rocks (ubiquitous term) 27, 203 deformability of rock masses 97
pre-Cambrian, Ko 299 deformation
upper crust 249 forced- apparatus 193, 194
Crystallaire well XTLR 90, 91 macro- testing 184
CSFT coupled stress/shear flow tests 453 micro- 184
shear-flow-gouge 488–492 modelled- 309, 310
tests with carbonate-equilibrated sea water 402, 487 moderate 304
tests with hot Ekofisk oil 401, 487 radial 304
tests with reservoir sandstones, chalk, shale 528–530 sub-millimeter size 309
tests with Sellafield ignimbrite/tuff 528–530 transient 242
CT-computed 196 deformation modulus (see also dynamic, Young’s, and
cube, cubic Qseis-similarity)
cubic specimens/samples 127, 137, 185, 303, 305, pseudo-static- (various symbols: Ed, D, M) 3, 5, 17, 20, 21,
359–364 46, 47, 49, 56, 92–95, 97–115, 108, 122, 134,
law 278 146, 162, 164, 165, 172, 175–177, 293, 300, 307,
rock volume, one fracture 472, 474 339, 348, 524
sandstone 419 as function of radial depth 301
structure 201 changes due to tunnelling 615
truncated 39 dependence on frequency (see dynamic moduli)
cubic network, idealized 172, 468, 480 339–341
cubically jointed 200 Ee (pseudo-static, rock mass ‘elastic’ unloading modulus)
Cundall’s distinct element codes 308–310, 365 97–103
curve-jumping, Q-jumping 93, 268, 272, 275, 285, gradient 107
286, 287 high- 22
cyclic rock properties increased 17
-effect on sonic log fluctuations 300, 301 inequality 97–115, 339
cylindrical specimens (ubiquitous term) low- 113
Cyprus 278 M-Q- ␴c relation 176, 366
668 Index

deformation modulus (see also dynamic, Young’s, and detecting point (receiver) 153
Qseis-similarity) (contd ) detonator cap 52
M-VP relation (note: M is pseudo-static estimate) 101, 167, deviated hole 299, 300, 305
365, 366 deviation (see rotation)
M-VP – ␴c relation 366 detector separations 9
M-VP –Qc relation 92, 94, 102, 115, 161, 167, 232, 257, Devonian sandstone-shale units 370–372, 370–371
258, 348, 365, 449 DHI direct hydrocarbon indicator 458
M-VP –Qc-L relation 164, 175, 176, 293, 365, 366 diabase 11, 12, 119, 121
mean- 114 unfractured 278
minimum- 114 diagenesis 206
reduced 5, 93 diameter
reduction 118 multi- influence, tunnel deformation 304
reduction due to extension fracture 362 diametral pair 307
ratio of Ee/Edyn (static elastic/dynamic) 107 diametrically opposite ␴␪max ␴␪min 306
ratio of Ed/Edyn (static deformation modulus/ diatomaceous earth 26
dynamic) 107 differential pressure, stress ( 1- 3)
tests 46, 50, 51 applied on model rock masses 524
tension fractured model 524 exceptionally low- 351
variable 227 exceptionally high- 539, 540
deformation modulus, similarity of high 538
static Eintact (when GPa units) to seismic Q 191, differential weathering
192, 355 diffusivity 464
static Ejointed (when GPa units) to seismic Q 200, 201, 362, dike, dikes 277–280, 264
363 basic 65
static Emass (when GPa units) to seismic Q 202, 203, 210, fractured, permeable 294
220, 221, 224–226, 269, 348, 350, 352, 365–367, mafic 215
387, 405, 410, 411, 424, 436, 476, 499, 500 sheeted 264, 278, 279, 281
e change in hydraulic aperture 489 dilatancy
E change in (mean) physical aperture 489 pre-earthquake 233
delta dilatant 303, 304, 307–310
-peak (␦peak) see joints non- 302, 303–305
weakly inclined- 302 strain 190
density 4, 5, 19 dilation 4, 84, 86, 169, 212
-compressive strength (shale) 26 causing build-up of stress 546
-depth trends (deeper, deeper holes) 111, 252, 253, 255, due to triaxial shear failure 540
264, 267, 298 enhanced- of fracture set 383
-depth trends (near-surface) 20, 78, 111 for enhanced permeability 435
-depth (sub-ocean sediments, Ontong-Java) 206 mobilized-, in stress transformation 553, 554
field- 21 onset of- 540
heterogeneities 191 peak- angles 525
high- anomalies 254–257 shear-induced- 390, 392, 406, 414, 444, 520, 522, 523,
of laboratory samples 19–23, 26, 263, 265 525, 526, 527, 531–533, 536, 540, 548, 550, 553
scaling, physical models 547 dilatation 5
VP data 19–23, 263–265, 267, 324, 326 diluvial clay 26
depth (ubiquitous term) 13 Diluvium 142
anomaly 14 diorite 20, 139–142
effects of- on seismic attributes 77–81 Äspö 139–142
estimation error 14 Äspö permeability 555, 556
increasing 13 Qseis 181
-modulus trends (see also deformation, and Qc ) 222 dipole
-pressure aspects of drilling 295–299 azimuthal, shear-wave, sonic logging 301, 302, 310,
to basement 8 312–316
to bedrock 9 dispersion cross-over 315
zone 19 flexural wave split into fast and slow directions 316
Index 669

modular- tool for radial, axial, azimuthal directions 316 dolerite


receiver array 314 intact 22, 23
receiver pairs 313 joint, stress-closure 494
tool orientation relative to formation 314 fissured 22, 23
tools pipe-conveyed for horizontal holes 315 joint, stress-deformation 495
transmitter tool 312, 314 joint normal stiffness, initial 495
transmitter pair 314 meta- 264
well-logs 382 weathered 22
directional permeability (see permeability, anisotropy dolomite 12, 32, 61, 100, 114
direct beds 486, 488, 520, 521
hydrocarbon indicator DHI 458 low porosity 31, 32
shear test DST 444, 452, 490–492, 510 higher porosity 359
shear test over-loaded, due to OC 517 high Qo 231
shear test, UDEC-model 518, 519 Ko 299
shear test, reconstruction of shear-dilation path 444 porosity – VP 325
wave 383 VP – density 326
discontinuity, discontinuities VP – VS 328
clay-filled Webatuck 31
different (micro-) scale 418 domain 173, 616, 622, 623
discovery and exploitation of new fields 396 domal structure
discrete Ekofisk 399, 401
particle model 203, 204 jointing 392
reflection 372 down-dip shear mechanism 401
dissolved air out of solution 133 down-hole
dispersion (frequency dependence) 182–184, 187, 189, 190, acoustic tool 73
191, 194 mud motor 312, 320
caused by microcracks or fractures, discussion 412, 413, 415 receivers and sources 378
caused by microcracks or fractures, modelling 460–481 seismometers 221–224
for reservoir characterization 387, 388 sensor array in overburden and reservoir 404
incorrectly assessed 379 sources, 4D time-lapse 352
intrinsic 387 VSP triaxial accelerometer sensor array 404
profiles 387 down-going
dispersive effect of fractures contra micro-cracks 5, multiple 383
412, 415 slab 179, 227, 228
displacement (of tunnel) 151 dragline excavators 118
discontinuity 309, 518–520 drainage and imbibition 337
modelled 309, 310 in presence of patchiness 335–337
to peak strength (see joints) 511, 523, 546 in presence of varied frequency 335–337
to peak, up-scaling 531–533 Draupne Formation, central Viking Graben, North Sea 352
discontinuity drift, drifts
clay-filled 156, 169, 188 multiple perimeter- 157
spacing 79 drill-holes (see boreholes, wells) 12
discontinuum drillability 17
lies on the floor 319, 320 drill-and-blast tunnelling 17, 30, 124, 125, 131–133,151
log-spiral fractured- 303, 305, 386 drillcore (see core) 11, 76, 77
modelling 400, 401 drilling (exploratory) 4, 9
partly connected 307 -induced fractures 73
discontinuity 189, 200 fluid pressure 299
displacement- events 189 mud 295, 296, 299
monitoring 134, 135, 151, 152 rate 150
distinct element model (see UDEC, numerical modelling) shallow crust from submersibles 266
Dixie Valley borehole data 543 super-deep 249, 252
Dneiper ship lock 117, 118 variable azimuth- (physical models) 302, 303, 305
‘dog-earing’ (see borehole breakout) while under stress 305
670 Index

dry 15 shear stiffness Ks dyn 423, 424


boreholes 73 stiffness testing 422–424
fracture/joint 197–199 stiffness modelling, UDEC 479
fractures, assumed effect on compliance ratio ZN /ZT 426 testing (ubiquitous term)
rock, perfectly dry 343 testing under normal load 197
state 5
state, contra saturated, joint stiffnesses 507–512 E and e apertures for joints or fractures (see hydraulic,
drying out 30, 31, 121, 353 aperture)
Drøbak, Oslo Fjord, Norway 56 Edyn Estat (see dynamic, see deformation modulus) 340,
DST (see direct shear test) 493, 507
dual porosity, see porosity E-modulus (see also Young’s, dynamic, and deformation)
ductile 37 191, 192, 424
-barrelling of sandstone samples 312 earth,
dynamic Core 182
bulk modulus K 13, 71, 104, 109–111 interior of the- 181
compliances (see compliance) internal structure of the- 52
cyclical events 373 – science’s favourite continuum theory 306
E-moduli (lab-scale) 201, 339, 340 earth’s crust
E-moduli, function of frequency 339 continental- 241–253
E-moduli, function of axial strain amplitude 339, 340 -continental margins 254–261
E-log 339 East-Pacific rise 273–287
E-low frequency 339 mid-Atlantic ridge 261–273
ejection, tunnel failure 304, 307 mid-Atlantic ridge and East-Pacific rise ages 287–290
field modulus (seismic) EF dyn 105–108 velocity structures 241–291
fracture compliances (see compliance) earthquake, earthquakes 52, 209–219, 257, 428–438
joint compliances (see compliance) attenuation (coda Qc1, QS1 and albedo: see attenuation,
joint stiffnesses (see normal stiffness) and Q section)
laboratory modulus (acoustic) EL dyn 108 attenuation of high-frequency energy 221
laboratory Young’s modulus (from VP and VS ultrasonic) back-scattering body waves from- 210
104 coda (see earthquake coda)
laboratory Young’s modulus (from VP and VS ultrasonic, as damage (caused by acceleration) 218, 219
function of stress) 126 distant 210, 211
loading 182, 422–424 deep 261
loading facility 192, 197, 357–359, 359–364, 422–424 deep- zone (sinking lithosphere) 227
micro-excursions 184 deeper heterogeneities above- 210
moduli 104–110 doublets 215
moduli M1,2 (-ratios at different frequency) 184 energy of 181
modulus Edyn (from VP and VS refraction seismic) 13, 71, epicentre 213
75, 107, 109, 110, 161 epicentral distance 216, 220, 221, 434
modulus Edyn (from AE interpretation) 130 event-distance effects 211
normal stiffness Kn dyn 200, 202, 423, 424, 478 focal depth 214
normal stiffness effects on velocity-frequency 478 focal depth change 212
parameters 7 frequency (see frequency, and Q section)
permeability discussion 425 geometric spreading 215
Poisson’s ratio (see Poisson’s ratio) hypocenter migration 215
properties 5 hypocentre relocation 257
properties of matrix 6 hypocentral distance 433, 434
properties of rock mass 6 hypocentral zone 210
ratio EL dyn/EF dyn (lab/field: dynamic) 106 local 210–213, 258, 428–434
ratio Edyn/Estat 340 magnitudes 213–219
ratio EF dyn/D (field: dynamic elastic/static deformation main shock 213
modulus) 107, 108 monitoring stations 211
ratio Ks dyn /Ks dyn 423, 424 pre-cursory changes 215
shear modulus ␮ 5, 13, 71, 104, 109, 111,161 QP and QS components 212
Index 671

rays from deep- 261 Misasa earthquake, Japan 213


reduced coda Qc after main shock 213 New Madrid seismic zone 428, 430
reduced coda Qc before main shock 213 Norris Lake Community, Georgia, USA 215
seismogenic depths 222 North Anatolian Fault Zone, Turkey 220
seismogram 209 Oishiyama earthquakes, Japan 210, 211
seismometers (three component) 219 Olkaria Geothermal Field, Kenyan Rift 394–396
seismometers (down-the-well) 221–224 Parkfield Dense Seismograph Array, USGS, California
seismometers (closely-spaced) 221–224 216, 221–224, 429, 431, 432
separation (intrinsic and scattering attenuation) Petatlan earthquake 213
213, 214 Rio Grande Rift earthquakes, New Mexico 210
shallow- 261 San Andreas fault zone SAFZ (90, 91, 95) 221–224
shallow heterogeneities above- 210 Stone Canyon earthquakes, California 210, 211
shallow-layer scattering and resonance 221 Southern Germany 220, 221
site effects 209, 220, 221, 224 Tangsham earthquake, China 110, 212, 213
small local- 212 Tsukuba Oishiyama earthquakes, Japan 210
spatial variation of amplitude, frequency-content, coda Vamanashi earthquake 213
duration 221 earthquake monitoring wells in California, specified
swarm 213, 215 Cajon Pass 221–224
take-off angles from source 215 Varian (array) 221–224
thrust-fault, shallow- 244 XTLR 91
travel-time data 258 earth-pressure-balance EPB 4
earthquake coda (see Q section) East Coast Margin Igneous Province ECMIP 254–256
late- 215 East Pacific rise crustal (see Q and Vp data lists)
low coda Q 215 velocity structures 273–289
high coda Q 215 East Texas well 441
spatial variation 215 EDA extensive dilation anisotropy due to cracks/micro-cracks
temporal variation during seismic events 210, 212–218 assumed source of shear wave splitting 409, 410, 412, 414,
earthquake, aftershock, locations, data, cited 428, 429, 432, 436, 453
Anza, California 222 critique 411–413, 470
Chi-Chi, Taiwan 436 Edgar Mine 64
Central California earthquakes 213, 214 EDGE 801 profile, Atlantic margin 256
Cerro Prieto Geothermal Field 435, 436 Edinburgh 405, 409
Chamoli earthquake, Garhwal Himalaya, India 218 Edinburgh Castle 118
Gazil earthquake, Uzbekistan 211 EDZ excavation damaged zone: 117–138, 151
group of locations for Qc-frequency data (Aleutians, EDZ excavation disturbed zone 51, 87, 93, 97, 106, 107,
Carolina, New England, Southern Norway, 115, 117–138
Canadian shield, Montenegro-Yugoslavia) 217 caused by joints, fractures, bedding planes 306,
group of locations for Qc -frequency data (Afghanistan, 308–311
Alaska, China, Eastern USA, Friuli: Italy, Guam, four-sector- 306–308
Hindu Kush, Iceland, Kinki, Southern California) measurements in tunnels 31, 119–133
228 multi-diameter- 308
Ghuj earthquake, Gujarat, India 219 of drill-and-blasted tunnels 119–133
Haicheng earthquake, China 212, 213 of shafts 119, 121, 123
Hawaiian earthquakes 213, 214 of slopes 117, 118
Húsavik, Iceland 432, 433 of TBM-driven tunnels 123, 306–310
Hyogoken Nanbu earthquake (and Tamba region), Japan of tunnels in general 301
217 velocities 140, 157
Iceland 416 EDZ, mini-, around boreholes 301–320
Iceland, Station BJA 433, 434 azimuthally varying- 301
Kuril-Kamchatka earthquakes 213 -discontinuum 302–306, 319
Loma Prieta earthquakes, California 215 -fluctuation in inter-beds 379
Long Valley earthquakes, California 213, 214 effective stress (ubiquitous term, see also stress) 263, 267,
Mamouth Lakes earthquakes, California 212, 213 295–297, 307, 351, 352
Mid-Atlantic Ridge axial valley site 436, 437 Egyptian oil-producing well 315
672 Index

Eikonal solver 16 England 11, 21, 29, 40, 42, 74, 79, 80, 94, 163
Ekofisk, Ekofisk reservoir 339, 392, 420, 438, 451, 487 North East 203, 207, 234, 346, 356, 379, 380
choice of deformation moduli at- 339, 443, 444 Southern- 319, 320, 488, 520, 521
choice of shear stiffnesses 427, 428 S.W- 488
gas-cloud difficulties 451, 499 enhanced
geomechanics-based 1D compaction model 399, 420, 443 oil recovery EOR 66
imaging through shallow gas 394, 396, 399 production techniques 363
index testing of conjugate joints 508, 510 environmental, environmentally
joint shearing 399, 401, 420, 443, 522 changes 30
modelling compaction at- with UDEC-BB 401, 443 effects 19
modelling subsidence at- with UDEC-MC 400 sensitive areas 172
monitoring with 4D 398, 399 Eocene 447
no discovery, no production if 549 Eocene limestone 11
pressure reduction 522 Eocene and Cretaceous, SW Texas 370
world’s first marine 3D/4C, 2002 451 EOR enhanced oil recovery 363
subsidence at- 339 gradient enhancement 295
elastic 4 permeability enhancement 295
acousto- 137 epicentre (see earthquake)
constants (static) 7, 183 epidote
constants (dynamic) 7, 183 mineral fillings 278
continuum theory 9 escalators 4
matrix 360 Ethiopia 245
moduli, modulus 4, 5 EU Hydratech Project 237
-parameters, time dependence in 4D 398 Eurasia 230
pseudo- response 306 Eurasia regions
reversible 183 Altaids 230
scattering 387 Arabian Shield 230
state 59 E. European Shield 230
stiffness matrix 417, 451, 477 Indian Shield 230
stress distribution, borehole 137, 302 Siberian Shield 230
stress distribution, tunnel 131 Ural Mountains 230
tensors, 3D 374 Eurasia, tectonic map, types 230, 246
waves (ubiquitous term) Shields 230
elasticity 373 Tethysides-Alpinotype 230
electrical Tethysides-Germanotype 230
circuit theory 181 European
conductivity 174 injection pressure limit (grouting) 162
resistivity (see resistivity) Eurotunnel 94
electro-magnet 187 excavation
electro-magnetic wave damage zone (see EDZ)
conductivity 169 effects 30
elliptical yielding zone 311 method
empirical (ubiquitous term) pre- velocities 165
equations (ubiquitous term) 256, 511 process 30
predictions 175 expansion due to high temperature 246
energy experimental testing facility (ubiquitous) 192, 197,
loss 18 357–359, 359–364, 192
loss/dissipated per cycle 18, 181 explosions
storage 18 nuclear tests as source 251–253
engineering exploratory drilling 4, 295
geological investigations 76, 469 exploration
geologists 58, 442, 469 borehole 306
-scale 179 -development well log display 316
units 161 galleries 99, 119
Index 673

infrastructure-led- 320 shear strength discussion, JRC, JCS possible 547


new- concept 372 throw magnitude, scale-dependency 447
on-land 4 traces from coherency analysis 392
marine transform-, intra-continental (NAFZ) 220
seismics 4 trap
explosive 63 valley-parallel fault swarm 154–156
charges 281 welded portions of- 427
source, sources 9, 243, 244, 254, 269, 389 zone 7, 9, 47, 48, 76, 79, 156, 157, 188
extension fractures zone compliances 427
created under biaxial stress 359, 361 zone model, Tod 462
rough- 364 zone, neighbourhood of 306, 308
VP VS 1/Q and permeability behaviour of- 361–364 zones, tunnel collapses, Q 558
extensive dilation anisotropy (see EDA) fault seals
extensometers 46 phyllosilicate 404
extrapolation of properties 49–51 cataclastic 404
extrusives (ocean-floor) 281 faulted
anticline 391
Factor-of-safety 306, 307 faulted rock 48, 161
zone rotation 306, 307 severely- 156
failure ubiquitous at all depths, KTB well 252
gradient, well pressure related 296 faulting
modes, tunnels 304 adverse- 363
surfaces, discretized beforehand 310 Fenton Hill borehole data 543
famous landmark 558 permeability 555
Faroe Islands 203. Fiji Islands 179, 227
fast filtering
and slow directions 374, 430 of high frequencies 199
and slow qS1 and qS2 directions 313, 314 filter paper (artificial fracture filling) 82
axis of formation 313 finely interlayered sequences 379, 380
compressional wave 209 first breaks 204
direction 38, 128, 276 First International Workshop on Scale Effects in Rock
Fourier transform 84 Masses 537, 538
polarizations 430, 437 fissures, water-filled 266
spreading ridge 281 finger-print, primeval 3D roughness 516
fault, faults, faulting 3, 65 finite difference (see numerical modelling)
active strike-slip- 257 Finnsjön, E. Sweden 203
causative 226 first arrivals 437, 479, 481
compliance 427, 501 fissure features 28
cemented- seals 404 fissured
domal 391 bulk samples 22, 23
either side of a- 314 non-fissured bulk samples 22, 23
first order transform- 283 rock 21
gouge 139 Fjellinjen road tunnel, Oslo, Norway 53
known- 24 fjord 56
major 164, 226 flat spots 458
major, VP ,L, Qc estimates 226, 556 Flatey Island, Iceland 433
mature sections of-, no dilation correction 554 flatjack loading (see also plate loading) 164–166,
minor 14, 164, 175, 510 513–516
minor, VP ,L, Qc estimates 226, 556 flexural shear waves 312, 314, 391
newer sections of-, dilation correction 554 polarized into fast and slow directions 312, 314
plane solutions 432 flow
regional fault zone 17, 47, 48, 53, 257 local 183
rosettes, water-flood directionality 403 macro- 183
swarm 411 no-flow criteria re fracture orientations 536, 541, 542
674 Index

flow (contd ) micro-scanner 379


simulator 398 undamaged- 313
versus fracture(joint) stiffness 425 foundations 3, 30, 224, 339
flow zone indicator (see FZI) bridge 19
fluid dam- 50, 51, 56, 159, 224
bearing lenses 414, 527, 550, 551 moduli 159
compressibility effects 397 pier 19
compressibility effect, oil or gas, ZN , slow S-wave powerhouse 172
407, 408 treatment (grouting) 170–172
-conducting joint structure 357 four-component 4C
crustal, lost with time 231 acquisition 394, 396
filled microcracks, preferentially oriented 408 sensing under gas clouds 394
flow monitoring, challenge 359 sensors 391
flow pathways parallel to ␴H max 526 four-dimensional 4D
flow rate 166 in-well installation 382
flow through local discontinuum 317 full-field 4D4C re-shoots 396
flow through log-spiral shear surfaces 317 frequent 4D4C surveys 396
-front monitoring, 4D 252 monitoring of reservoirs 396, 397
-lense rotation phenomenon 390 monitoring of compaction 398, 399
loss 231 need for velocity and attenuation monitoring in 4D 397
-past-frame attenuation 356 repeated surveys 372
path assisted by joint roughness 358 seismic in fractured reservoir 364
pressure close to ␴H max , critique 416 time-lapse monitoring 352, 353
micro- 282 time shifts, in compacting reservoir 398, 399
saturation 8 frac gradient (see fracture) 296
saturation effects 183 fractal
type from compliances, theories 425, 426 characteristics 223
velocity of- 30 earth 179
viscosity 457 re vertical wells, vertical structure 441
fluid substitution theory 407 fracture, fractures (see also joints, cracks, tension-)
contrary to the wisdom of- 407 alignments, alternative to ␴H max 410
flysch sandstone 25 aperture (see aperture)
FMS log, analysis 389, 455, 525 azimuth 382–386, 388–396, 401–406, 407–482
Foidel Creek, Colorado 60 behaviour differences, rough and smooth 363
foliation(ubiquitous term) closure cycles 363
joint 513, 514 closure-velocity stronger result 363, 364
planes 355 closure-permeability weaker 363, 364
Folkstone Warren, S.England 320 compliance (see compliance)
formation compliance tensor 479
density tool (compensated) 47 compliant- 407
fast axis 409 conductivity (see permeability)
fast axis substitution for ␴H max 409 criticality (Crampin) 413, 434
micro scanner 47 densities 224
slow axis 409 density:N/V.r3 407
fossil-spreading direction (sub-ocean crust) 276 density, inversion for 471, 474
fossil density, need for volume-defined 407, 455
flow permeabilities 278 density in reservoir, model 472, 474
flow porosities 278 density variation between two sets 404
spreading direction 276 detection 377
forced deformation apparatus 193, 194 dilation, model 444, 519
Forgario: limestones in situ 105 displacement 87
formation fluid-flow pathways 373
damage (see borehole, see EDZ) fluid investigations 314
evaluation 312–319 frequency variation at fault zones 394
Index 675

gradient 296 parallel to ␴H max measurements 402, 403, 447, 448,


heavily-, numerical (UDEC-MC) model 308 450
heavily-, physical model 524 rough fractures, hard rock needed for parallel ␴H max 541
induced-, high pressure triaxial, shear strength 537 fractured
investigations 313 badly 163
like cockroaches (families, all scales) 373 heavily- 7
liquid-filled rather than gas-filled- 391 heavily-, generic case 309
mechanics code (FRACOD) 306, 307 moderately 11
network 172 reservoirs 179, 369–406, 438–459
normal or -strike anomaly, AVO 391 reservoirs, ‘open’ fractures/joints, tests, models 528–534
open- (see next) petroleum reservoirs 3, 35, 373
O/R open, rock-to-rock contrary rotations 414, 446, 450 petroleum reservoirs, extend current thinking 483
orientations from outcrops, FMS, core 250, 252, 382, 385, rock mass 5
472, 475, 525 sample 197–200
orientations (from AVO) 390 zone 73, 153
pattern, outcrop 448 zone width 73
permeability and stiffness link 425 fracturing
permeability, changes of 372 anisotropically distributed 304
permeability, exploring for 372 before-and-after- VP VS Q1 and permeability 362–364
permeability tensor 479, 480 better identification of-, with shear-wave technology 407
permeability-velocity-loading hysteresis 363 development over time 306
porosity 457 due to over-pressure 296, 352
radius versus density 472, 474 pervasive 273
rosettes 385, 403, 404, 406, 441, 448, 475 stress induced- 314
rosettes, water-flood directionality 403 stress-related (see borehole, tunnel) 302
rosettes, reservoir and overburden 406 tectonic 273
rough extension- 363, 406 thermal
rough extension- showing ‘open’/‘closed’ rotation 406 Franciscan rocks, California 222, 257
sealed- parallel ␴H max are numerous 441 QP and QS 222
sets, two, perpendicular 447, 448 France 39, 76, 185, 320
several- systems (in AVO case) 390 freeze 32
shear and normal compliance ratio 393 -thaw cycles 185
spacing (Fm1) 11, 57, 69, 70–72 freezing 32, 33, 53, 56, 185
stiffness (see stiffness, joint stiffness) frequency 112, 183, 198
stress alignment excellent, near-surface 447, 448 and viscosity (product of ) 207
strike, dominant 383, 389 band 182, 184
strike, production trend mismatch 383 dependent 5, 17, 183
swarms, below seismic detection 455 dependent cross-well data 378
tensile- traces, from point load 385 dependent energy loss 190
tension- , physical models 523, 524, 551, 552 domain 118
zone 11 dominant- 7
zone, mid-ocean ridge 273 dynamic normal stiffness, joints 498
fracture characterization from effect on VP with patchiness 335–337
core 382–386, 389, 404–406, 446, 447 high-, low- differences re scattering, intrinsic attenuation
borehole logs 382–386, 389, 404–406, 446, 447 387
outcrop analogues 382–386, 389, 446, 447 high- propagation over long distance 378, 478
fractures, open high- P-wave monitoring 301
conducting- 404 higher-, shallower viewing LWD velocities 318, 319
conducting- due to shear and dilation 406 independence below 1 Hz assumption 232
critically (shear) stressed crust 541, 542 influence on attenuation in jointed, bedded rock
due to shear strength, dilation, in situ block sizes 207–209
534, 535 instantaneous-, difference versus depth 459
from 0° to 90° from ␴H max 441 low-, for imaging, sub-basalt 403
parallel to ␴H max assumption 390 lower-, deeper viewing wireline velocities 318, 319
676 Index

frequency (contd ) link to RQI reservoir quality index 333


negligible- dependence 239 logging-based- 336
power laws relating seismic Q and f 205, 212, 217–219, relation to permeability 333–336
228 relation to porosity 333–336
predominant- 212 sorted by- 334, 335
frequency dependence of
coda Qc 210–212, 214, 217–219 Gabbro 11, 12
damage from freeze-thaw 185 meta- 264
Edynamic 339, 340 olivine- 264
E moduli (97–105), 339, 340 pyroxene- 264
fluid mobility 194 unfractured 278
magnitude spectra, jointed samples 199 gallery
Q, Qseis, QP 187, 189, 194, 207 between- seismic 49–51
QE 190 -to-gallery 49–51
1/Q of intact samples 200, 343 gas
1/Q of jointed samples 200 and condensate 343
Q, 1/Q in situ 209, 238, 239 bearing sandstone reservoirs 343
1000/Q intact samples 191, 344 brine, distinguishing 455–458
1000/Q in situ 208 chimney 456
1/Qc (coda Q,) from earthquakes 210, 211, 214, cloud effect 399
217–219, 223, 228 cloud penetration 396
1/Qe,p,k,s 194 condensate 343
strain amplitudes 182–184 detecting-, (when over-pressure) 343, 349, (350)
VP intact samples (patchy saturation) 337 extraction (earthquake source) 211
VP intact samples 343, 344 filled fractures, compliance ratio discussion 393, 417–421
VP jointed samples) 196 fractured- reservoir 471–475
VP in situ (e.g. sonic/VSP) 207, 208, 234, 252, 280, generation 296
313–315, 318 hydrate sediments 237, 238
friable rock classes 28 oil contact
friction (cracks, joints, fractures) outbursts 60
basic- angle (␸b) 509, 510, 538 pockets squeezing 182
basic- angle, effect of moisture 511 producing natural fractures 384
coefficient 183, 535, 540, 542, 543, 540, 542–544 sand 15
crack-tip- 196 shallow- 407
critical state- angle, ␸c 538, 539, 544, 547 time-sag due to- 381
grain boundaries 183 zone in well, QP and QS separation 345
mobilized- , conducting fractures 542–545 Gassmann
macroscopic- 188 fluid substitution 458
non-linear 17 predicted 328
opposed by 182 theory for porous rock 421
residual- angle (␸r) 509–512, 522, 525, 526, 531–535, theory error for fractured porous rock 421, 458
539, 547, 548 garnet-free 256
residual- angle, effect of moisture 502, 509–512 Geodia 110
frictional geology 4, 9, 255
attenuation mechanisms (see attenuation) steeply-folded- 408
dissipation 182 structural 9, 214
mobilization (see JRCmob) geologist
mobilization at larger strain 304 engineering- 11
sliding at grain contacts, boundaries 188, 219, 220 geological
fused glass beads investigations 19
attenuation – frequency 345 profile 9
lack of pressure dependence 343 reality 14, 15
VP – frequency 345 structures, major 65, 169
FZI flow zone indicator 333–336 time/age 370–371
Index 677

geomechanics 179 measurement locations 246, 251


school of thought 534 new- model 249
geometrical spreading 280 optimization 445
geophone 7 review of crustal velocity structures 244–251
in-line, vertical, cross-line 391, 394, 396 Glomar Challenge drilling ship 263
position (VSP) 383 gneiss 4, 11, 12, 13, 32, 38, 100, 139–142, 222
spreads 42 CSFT, gouge formation 492
geophysical 3 deeply weathered 4
logging 153 foliated 38
team 9, 169 ‘geologically uniform’ 215
geophysicist, geophysicists 17, 95, 219, 369, 407, 416, 421, in situ 527
469 jointed (in situ) 16, 70, 88
posing questions to- theoreticians 520, 531 Ko 299
Geophysics (journal) 378 Mojavi, S. California 261
geophysics 98, 197 QP and QS, Cajon Pass drillhole 222
literature 6, 488 schistose- (EDZ) 123
principles of 434 sequences (KTB) 249
geophysics jargon without explanation tolerance of stress anisotropy 298
CMP-gather 234 tunnel EDZ in- 121, 122
CMP number 235 weathered 107, 108
NMO-corrected 234 god-given means 407
QVO quality versus offset 234 Gonwanaland 231
true-relative-spectrum-processed 234 gouge production (see joint)
Geoteam 148 GPS clock 153
geotechnical graben 238
impossibility(?) 415, 416 grain, grains
investigations 19 boundaries 182, 183
geothermal sites boundary cracks, due to stress unloading 397
Beaver County, Utah 42–44 boundary sliding, attenuation 219
Cambourne School of Mines, Cornwall 169 boundary weathering 264
Cerro Prieto- Field 435, 436 contact microcacks 347
Olkaria Field, Kenyan Rift 394–396 density 343
steam-dominated –field, lower Poisson’s ratio 394 loosely-packed- 206
water-dominated –field 394 size 343
geothermal size squirt-flow, no fractures 475
gradient 351 granites 4, 11, 12, 13, 61, 100, 139–142, 196, 222
-metamorphic depth-permeability curve 555 aspect ratio influences 326
geotomography (see tomography) Äspö 131–133
Germany 249, 252, 543 Äspö permeability 555, 556
Southern (seismicity) 220, 221 artificial surfaces 80, 82
Gezhouba hydroelectric project, Yangtze River, China Barré 36
172 basement- , Egypt, fault zone 315
gilsonite dikes 472, 475 Cajon Pass borehole 541, 542
Gjøvik (Olympic) cavern, Norway 16, 43, 53, 55, 74, Casco 31
88, 89 Chelmsford 30
Ghuj earthquake, Gujarat, India 219 Cornwall (in situ) 40
glacial decomposed 11
post- landslide 95 deeply weathered 4
till 12 in situ, block test 527
glacier intact 11
glass (see also fused) jointed (in situ) 11, 24, 40, 41, 70, 72, 159, 165, 166
beads, low Q 205 Ko 299
value of Qseis 181 massive 128–131
global plutons 243
678 Index

granites (contd ) Haematite


QP and QS, Cajon Pass drillhole 222 Haicheng earthquake, China 212, 213
1/Q (1/QP) – axial stress-strain to failure 196 Hales method 15
Qseis 182 Haltenbanken, Norwegian sector of North Sea 312
shear strength in situ 299 hammer seismic 3
Sierra white QE Hanford, Washington 119, 124–127
Stripa (quartz monzonite) 133–136, 422–424, 435 hard (brittle) rock (ubiquitous term) 6, 7, 13, 74, 76, 130,
stress-conducting aperture, CSTF 529, 530 147, 272
Tertiary 19, 21 beds (masking) 146
tolerance of stress anisotropy 298, 299 crystalline rocks (ubiquitous term) 210, 527
Troy 31 jointed rock Qc Emass VP 226, 556
tunnel EDZ in- 121, 122 massive rock, Qc Emass VP 226, 556
URL Manitoba 128–131 rock sites 12, 70–72, 160
weathered 20, 21, 24, 77, 107, 108 hardness 27
Westerley 31 harmonic excitation 181, 205
granodiorite 7, 8, 12, 72, 100, 203 Hawaii 213, 214
artificial surfaces 82 Hawaiian
Cajon Pass borehole 541, 542 earthquakes 213, 214
Ko 299 Islands 257
QP and Q S , Cajon Pass drillhole 222 ridge 257
granulite to ecologite transformation 242 head
gravel 11, 20 -race tunnels 147, 154–157
formation 205 waves 16
gravity heated
anomalies 255–257 block test 513–519
Great Lakes, Canada, USA 231 flow tests 515, 527
Green River (Upper/Lower) reservoir formations joint samples 197
471–475 mine-by 135
greenschist heater-borehole 134–136
sub-ocean floor 281 heat-flow assumptions (crustal) 248, 249, 555
greenstone 32 heat-producing concrete 130
GRM generalized reciprocal method 15 heavy oil (194)
Grimsel, Switzerland 124, 169 heavily jointed rock mass 9
groundwater Helsinki 72–73
pressure drawdown 53 Hertz contact theory 23
groundlevel 3 Heterogeneous Earth 213
grout heterogeneity 373
curtain 57 heterogeneous
monitoring (with velocity) 170–172 conditions 4
penetration/take 172, 490 re-classified as severely- 373
prediction 479, 487, 490 severely 373
grouting (see also pre-injection) 88, 89, 170 hexagonal-crack, APE model 415
after- 170–175, 292, 293 high
before- 170–175, 292, 293 explosive (see explosive)
efficiency 171 fluid pressures 320, 321
Guadalupe Island 267 frequency (see frequency)
guillotine, double-bladed 523, 524, 551, 552 pressure high temperature well (see HPHT) 320, 321
Gullfaks field, North Sea 398 pressure (ubiquitous term)
Gulf Coast reservoirs 311 resolution images (ubiquitous term) 3, 379
Gulf Coast Plain 231 -rise building collapses 219
gypsum samples -speed layer 14
CaCO3 and water, cement component 547, 548, 551, strength rocks (see also hard rock) 101, 103
552 stresses (see stress) 320, 321
with flaws 40, 41 temperature 248–250, 320, 321
Index 679

velocity layer 49 fractures (as reflectors) 249, 252


voltage sparker 49 fracturing (minifrac: stress measurement) 19, 53, 89,
highway cutting 25 298–300, 406, 536
Hijori deep borehole data 555 fracturing, natural, recent 472
Himalayan dam sites 12, 122 head (ubiquitous term)
history matching jacking (of joints or major discontinuities) 162, 163
to parameter estimation to prediction of performance 396 modellers 133
Hole 418A, Eastern Pacific 276, 279 test/testing 135,161, 174
Hole 504B, eastern equatorial Pacific, Costa Rica Ridge theory 173
area 276, 277–280 units 333
Hole 648B, mid-Atlantic ridge 267 units sorted by FZI 333–335
Hole 735B 279 hydraulically
Hole RCF-1 (Sellafield) 387 conductive fractures 541, 542
hole non-conductive fractures 541, 542
collapse 320 hydrocarbon, hydrocarbons
stabilization 305 -bearing rocks 295
homogeneous bearing rock, neither isotropic or homogeneous 382
-and isotropic -bearing structures 369
-and isotropic, or heterogeneous and anisotropic 382 contact 297
isotropic elastic 382 escape to atmosphere 295
Hong Kong 74, 76 exploration 295
Honshu-Shikoku Bridges 19, 21 migration 295
Homestake gold mine 60 production, significance of vertical structure 407
homogenous zones (tunnel) 149 prospect 14
horizontal hydrologically-predicted pathway 377
drilling 315 hydro-mechanical testing 359–364
stress (see stress) hydropower, hydroelectric projects (see also dams, dam sites)
horizontally bedded (ubiquitous term) 374 tunnels 120, 121, 147, 154–157, 171
hornblendite 32 hydrophone
hornfels 38 detectors 62
host rocks 22 receiver array 67
hot dry rock project (see also geothermal) 195 receiver, ocean-floor 267, 269, 394
HPHT hydrostatic
American unit classification 320 gradient 296, 297
extreme- 320 normal- pressure 297
high pressure high temperature 320, 321 pressure 296, 297
Tier I, II, III 320 hydrostone, roughness profiles 509
three-tier classification 320 hydro-thermal
ultra- 320 alteration link to velocity 287
HRSN (see Parkfield) and mechanical hysteresis and over-closure 513
HTI (TIH) media 382, 394 deep- fluids 231
HTM modelling deposition 270
effects during water-flooding 401 fluid injection 270, 278, 279, 293, 294
fully-coupled- of water drive in jointed rock 402 fluid circulation 285
Hudson diagonal (compliance scale-effect model) 503, hydro-thermo-mechanical
504, 515, 526 in situ block test 513–517
Húsavik, Iceland 432, 433 detail, not in models yet 517
Húsavik-Flatey fault, Iceland 433 hydrotomography 174, 442
hybrid machine 4 Hyogoken Nanbu earthquake (and Tamba region), Japan
hydraulic 217
aperture (see aperture) 83, 161–163, 169, 198 hypocentre (see earthquake)
apertures in fracture zone 506 hypothetical (ubiquitous term)
conductivity 3, 162, 163 mineralization 336
diffusivity 169 pore-filling 336
680 Index

Hypabyssal rocks 142 inelastic


hysteresis behaviour
affecting velocity-permeability coupling 364 inelasticity of matrix, intrinsic attenuation 387
on first-closure cycle 87, 364 inequality of
Høvik, Norway, quartz lense with shear 551 deformation moduli 97–115, 339
Edynamic  Estatic 104–108, 184, 339, 340
Ice 12, 249, 253, 254 dynamic fracture compliance  (static fracture stiffness)1 184
fern 253 inertial
formation 32, 253 effects 207
-front 32 forces 182
glacial- 253, 254 infinity (") 182
occupied by- 33 Inguri arch dam, Georgia 88–90, 102, 170, 171
-wedging 33 input signal 118
Iceland 228, 416, 432–434 injection (see pre-injection, grouting)
Idaho mine, Idaho 321 in situ (ubiquitous term) 339
Idaho Springs, Colorado 64 block test, BB-simulation 507
IFP Picrocol data set 381 deformability 19
igneous rocks 6, 12, 25, 71, 72, 105, 109, 246, 263, 278 natural block size 507
massive- 163 P-wave velocities 40
fractured-, permeability range 279 scaling of stiffness(shear) 421, 422
garnet-free 256, 257 strength 19
seismic Q 186 Institut Français du Pétrole 357–359
unweathered 13, 72 instrumentation 165
ignimbrite 94 integrated model 92
stress-conducting aperture, CSTF 529, 530 interbedded, interbedding (ubiquitous term) 296–300
Ikeda rock classes, Japan 142 brittle layers 297–299
IKU, Trondheim 470, 480, 481 plastic layers 297–300
illite 27, 351 reservoir-type rocks (ubiquitous term) 356
impedence sandstone and shale 207–209, 300, 301
-changes in 4D 252 sandstone and shale, numerical model 310
-contrasts 202, 280 shale/siltstone, Kimmeridge Bay 520, 521
for distinguishing 326, 369 units 369–372
Imperial College, London University 517 interfaces
polyaxial cell 359–364 shallow 7
Borehole Test Site, NE England 207–209, 234, 346, 356, integrated effects 19
379 internal friction (dissipation factor) 205
Imperial Valley, southern California 260 inter-related properties 21, 25
impermeable 555 intrinsic absorption (see attenuation)
important insight intrusion
coupled earth-science behaviour 197 dike- 277–279
incident angle variations 354, 355, 466, 470, 471, 480, intrusives
481 high MgO 255
with off-vertical fractures 419, 420 mafic 255
with off-vertical- 419, 420 invaded zone (see mud infiltration) 316–318
independent constants 374, 477 investigation galleries 47, 48
index tests (see joint index tests) ionic effects (of clay) 169
India 218, 219 Iran 107, 108
N. 302 ISONIC sonic-while-drilling tool 313
SE coast of- 219 isothermal (static) 183
Indian isotropic medium, stiffness matrix 374, 477
Peninsula 219 isotropic elastic continuum analysis 305
Shield 219, 230 jointed reality not modelled 401
industry assumption, standard 390 mismatch with discontinuum model 400
regarding permeability, fractures, ␴H max 390 mismatch with subsidence bowl 400
Index 681

Israel 54 formation temperature, re thermal OC 516


ISRM congress 97 frequency (F m1, ␭ m1) 12, 13, 16, 17, 41, 43, 69, 77,
ISRM commission on failure modes 302, 303 81, 82, 83, 84, 85, 88, 89, 150
ISRM (1978) describing discontinuities, terminology 419, gouge formation with shear 489, 491, 492, 526
469, 511 hysteresis, of fractures 87, 364
Italy 60, 77, 101, 102, 171, 228 index tests (see joint index tests)
intersecting- affecting stress-closure 364
Jack-up of platforms, Ekofisk 442 joint-or-fracture terminology, ISRM 419
Japan 19, 21, 40, 41, 65, 94, 101, 104, 148, mated and unmated behaviour 87, 494, 506
149, 195, 210, 211, 213, 217, 228, 229, 243, model (UDEC) contact apertures, lengths, angles 521
244, 304 modelling ignored, when too large scale 401
N.E.- arc 243 near-vertical (see sub-vertical)
Sea 245 non-conducting 172
Trench 245 non-planar-, counter rotation effect 384
Japanese non-vertical- bias, vertical holes 385
authorities 19 normal stiffness 16, 45
data 21, 25, 27 number of- 30
Highways classification 21 numerically-glued 133
high-speed railway tunnels 47, 48 open (see fractures-open, stress H max, non-aligned)
rock classes (Ikeda) 142 O/R open / rock-to-rock, contrary rotations of shearing-
Java (-Ontong Platform) 206 414, 446, 450, 453
Jernbaneverket, Norwegian Rail Authority 173 orientation 40–45
joint, jointed, jointing, fracture, fractured (ubiquitous, and permeability-shear 491, 492
interchanged terms) pole concentration 89
apertures (see aperture) 17, 80, 82, 84 primary- 414, 548, 551, 552
apertures, intra-bed 208 primary cross- 551, 552
apertures E and e, JRC related 453 related overbreak in boreholes 250, 252
aperture, unstressed, equation 511 replica roughness profiles, ‘large’ scale 545
artificial saw-cut- 77 rosettes 385, 403, 404, 406, 441, 448, 475
bed-limited 208 rosettes, water-flood directionality 403
cleavage 490, 491 rosettes, reservoir and overburden 406
closure 13, 19, 81–86, 135, 136, 199, 258 roughness 11, 83, 84, 85, 169, 197, 422, 451
closure effects 285 roughness coefficient (see JRC)
connectivity needs at geothermal site 436 roughness coefficient (JRC) profiles, lab, large-scale lab 508,
conjugate 88, 401–403 509, 544
conjugate- at Ekofisk 401, 452 roughness contribution to strength, large scale 544, 545
compliance (see compliance) roughness mobilized ( JRCmob) 489–491, 502
component sets of- 44, 45 roughness destroyed 491, 502
cooling 269 sample, static, dynamic, flow testing 197–200
core recovered from Ekofisk 401 saturated or dry stiffnesses 507–512
cross-column jointing (basalt) 124–127 secondary- 414, 550–552
deformation 161, 162 sensitivity to- 65
delta peak (␦peak) 491 set alignments, with ␴H max as bisector 483
density 224 sets, more than one (ubiquitous) 483
dilation (see dilation) 490, 491 sets, primary, secondary 523, 524, 551, 552
displacement 166 shear magnitude 308, 433
displacement to peak strength 507, 511, 522, 523, 546, shearing 110, 165, 299
548 shearing most where high porosity 402, 433
displacement discontinuity 518–521 shearing, modelled 309, 310, 443
dry or saturated stiffnesses 507–512 shearing, physically modelled 547, 548
due to anticlinal structure 381 shear stiffness (see stiffness, shear stiffness)
effect on attenuation 188 slickensided, due to compaction, conjugate shear 443,
effect on Qseis 197–200 452
filler, filling 30, 84, 85 spacing 139, 469, 472
682 Index

joint, jointed, jointing, fracture, fractured (ubiquitous, and RQD/Jn as relative block size (see also scattering) 225, 469,
interchanged terms) (contd ) 615
sparse 307 Jr -joint roughness rating 6, 53, 92, 144, 175, 225, 293,
sparsely-, massive QH2O – K estimation 557 350, 447, 448
stiffness (static/pseudo-static,) 5, 6, 45, 79, 98, 106, 184, Jr/Ja as friction coefficient ␮ for joints, fractures, faults 225,
418, 420, 463 350, 411, 469, 615, 620
stiffness changes 309 Jr/Ja as friction coefficient of clay-filled discontinuities 538,
stiffness, dynamic, effects on velocity-frequency 478 556, 615, 620
stiffness non-linearity 420 Jr/Ja as directional parameter causing anisotropy 411
steeply-dipping (ubiquitous term) 357, 401 Ja/Jr for Qwater and for QH2O 556–558
strength displacement dilation BB-modelling 506, 507 Jw -joint water inflow rating 6, 92, 169, 175, 225, 293,
strength, limited 308 350,447, 448
stress discontinuity 518–521 (see SRF for 6th Q-parameter)
stretch in subsidence bowl 412 J-parameters from Barton, Barton-Bandis joint model
strike, rotating in domal structure 401 JCS joint wall compressive strength 79, 83, 203, 258, 359,
structure effects 285 420, 422, 424, 425, 428, 451, 456, 485, 492, 495,
sub-vertical 40, 41, 42, 359 496, 507–512, 540
surface 9 block test 518
surface area, reason for water weakening 402 JCS and moisture 502, 507, 509, 511, 512
testing 282 JCSo 526, 543–549
testing, stress-closure 282 JCSo in CSFT 529, 530
testing, shear-displacement-dilation 282 JCSo to JCSn scaling 531–535
three sets of- 385 JCSn 523, 525, 541, 543–549
vertical (see sub-vertical) ratio JCS/␴n 79, 258, 485, 508, 520, 547
void 119 ratio ␴n/JCS in CSFT tests 528–530
wall compressive strength (see JCS) scaling equation 511
water-conducting- 172, 173 scaling with block size 523, 525
weathered, partly 508 JRC joint roughness coefficient 83, 84, 86,162, 197, 359,
weathered, SW, MW joint stiffnesses 496 363, 364, 420, 422, 424, 425, 428, 451, 456, 457,
jointed 304 460, 467, 485, 496, 507–512, 540
heavily- 7, 163 back-calculation of- from tilt test 509, 510
less frequently- 19 block test 518
-rock 30 JRC ‘at right-angles’ 517
-rock masses (ubiquitous term) 4, 5, 124–126, scaling equation 511, 544
365, 367 scaling with block size 523, 525
-rock masses, coal analogue 168 JRCo 519, 526, 543
-rock model JRCo importance in strength-deformation scaling 535
sparsely- 7 JRCo in CSFT 529, 530
joint index tests JRCo and JCSo 487, 493, 508, 510, 543–549
Schmidt hammer 26, 486, 510, 545 JRCo to JRCn scaling 531–535
tilt tests 401, 508, 510, 513, 514, 545 JRCn 523, 525, 541, 543–549
tilt tests, large-scale 509, 553 JRCn estimate from tilt tests 513, 514
wall compressive strength (JCS) 425, 508, 510, 545 JRCn measured in DST and biaxial shear tests 523, 524
wall roughness(JRC) 401, 425, 510, 545 scaling- to block-scale JRCn and JCSn 484, 508, 510
joint industry project 303 JRCmob concept 489–491, 507, 545
Juan de Fuca Ridge 269, 279 applied to model tension fractures 548
Jurassic 185 applied to stress transformation, corrected 553, 554
sandstone-shale units 371
J-parameters from Q-system (see Appendix A for all numerical Kamioka Mine 65
ratings, descriptions 615–626) Kane, North Atlantic fracture zone profile 273
Ja -joint alteration, clay-filling rating 6, 53, 92, 144, 169, Ko (ratio of ␴h min/␴v) 299, 300, 443
175, 225, 293, 350, 447, 448 increased by joint shear in 1D compaction 443
Jn -number of joint sets rating 6, 53, 92, 144, 169, 175, reversal at shallow depth 299, 300
202, 225, 250, 293, 350, 447, 448, 469 kaolinite gouge (artificial filling) 83
Index 683

karst, karsts 4, 196, 197 lattice crystal structures 509


karstic voids 172 models 265
Lawrence Berkeley Laboratory, California 197–200 sequence 233, 234
Kern River oil sand 330 velocities and densities (sub-ocean crust) 267
kerogen layer thickness
content 351, 353 estimates 14
-to oil conversion 352 minimum detectable- 379
Kielder aquaduct tunnel 79, 80 reduction during compaction 398–402
kinematic, post-kinematic (see aperture, sealing) Layer 1, 2, 3 sub-ocean divisions 243, 244, 255, 257
Kimmeridge Bay, Dorset, S. England 488, 520, 521 Layer 2A 267, 268, 283, 286, 287–289
King, Imperial College polyaxial cell 361, 365 Layer 2B 278, 283, 287
kink band in physical model 524 Layer 2A, 2B, (2C) 265, 267, 271, 272, 277, 283–285,
Kitakami massif, Northern Honshu, Japan 229 287
Kn and Ks normal and shear stiffnesses (see normal, shear Layer 4, 5, 6, sub-ocean divisions 244
stiffness) lead, Qseis 181
Knopoff ’s seismic Q of selected materials 181 Lg coda (cross-continent seismic Q) 230, 231
knee-shaped velocity-depth curves 241, 243, 245, 250, Liaoning earthquake, China 110
252–254, 260, 261, 262, 264, 266, 267, 271, Lias shale 11
274–277, 280, 281, 284–286, 289, 290, 311 life of field seismic LOFS (453)
Koefels landslide 95 lime quarry 61
Kola Peninsula super-deep well data 555 limestone, limestones 4, 12, 32, 100, 108, 114, 370, 385
KTB super-deep well 249, 250, 252, 542, 543 aspect ratio influences 326
permeabilities 554, 555 basic friction angles, dry, wet 511
Kuril-Kamchatka earthquakes 213 Bedford 31
Kurobe IV dam site 40 Carboniferous 207–209, 346–348, 356, 357
Kurtachov, North Atlantic fracture zone profile 273 CBTF well- VP – VS – density-permeability-birefringence
445, 446
La Cira-Infantas Oil Field LCI 332, 334 Corralian 79
laboratory (see VP and Qseis listings) crystalline 12, 101, 102
high pressure tests 22, 31, 58, 59, 248–250, 256, 257, density-VP 325
323–364 hard 12
samples (ubiquitous term) 259, 323–364, 339 finely interlayered 379, 450
test, large-scale 352, 353 foundations (dams) 49–51, 97, 98
test and large-scale BB-modelling 507 fractured, chalky-lst. 455–458
tests on sub-ocean basalts 261, 263 fractured, well 297, 389
triaxial tests 258 in situ 42, 43, 45, 105
ultrasonic velocity 70, 73, 91, 183, 207–209 intensely jointed 102–104
Lagerdorf chalk quarry, Germany 420 interbedded 47, 207–209, 212, 234
Lamé’s constants 373, 374, 417 joint, initial stiffness 494
laminar flow 161, 474, 513 joint normal stiffness – stress 496
laminations, shale 355 joint normal stiffness, initial 495
Lamont-Doherty Geological Observatory 261 joint, stress-closure 494
landfill 7 joint, stress-deformation 495
large scale rock properties 165–167 jointed 350, 455–458
laser scanner micrometer (non-contacting) 84 Jurassic, bedded- 185
lateral expansion (see Poisson’s ratio) Jurassic and carboniferous, compliance, saw-cuts
Latiyan dam site, Iran 107, 108 501–503
Lau Basin 179, 227 Ko 299
Lau Ridge 179, 227 low porosity 31, 32
Laupies dam site, France 159 marly 101, 102
layer, layered, layering medium jointed 102–104
-cake sequence nodular 148
fine horizontal- 357 Oncolithic, Grain-, Pack-, Wackestone 349
inter-beds 3, 45–47, 207–209, 234 Oolitic, water, oil saturated 348, 349
684 Index

limestone, limestones (contd ) shearing 302–307


Permeability extremely high 378 Loma Prieta earthquakes, California 215
porous 24, 31, 32 London-Brabant Massif, North Sea 447
porosity – VP 325 Long Valley earthquakes, California 213, 214
Qseis (QP) 181, 186 Long Valley well data 542, 543
reduced (␣-coefficient) 321 long wall
-shale interbeds 148 mining 53, 61
soft 12 pillar 60, 61
Solenhofen 31 shearer 53
stress-closure tests on joints 528 Lorraine, France 185
thin-bedded (EDZ) 121 Los Angeles Basin, southern California 260
Triassic, weathered and jointed 98 lost circulation 299
tunnel EDZ in- 119 low velocity
Vajon 108 layer 14
VP – density 326 surface sediments 231
VP – VS 328 zones 71
1000/QE and 1000/QP confining pressure, crinoidal Lower
191 Cretaceous 391, 447
Lincolnshire 42 Hod formation 391
line mantle 241
-drilling (tunnel excavation) 128, 130 Paleozoic 447
-drilling (of flatjack slots) 513 Permian 447
linear Silesian coal basin 60
elastic isotropic medium 373 Lucite
regression 256, 257 honey-saturated- plates 427
liquid- or gas- filled fractures 391 laminate model, critique re ZN and ZT ‘equality’
Lista Formation 391 417–419
lithological units QE -value 188
causing multi-stepped response 336, 337 Lugeon test
mixed 335 apertures e and E for grouting design 172–176, 468
mixed causing patchy saturation 335–337 conversion to permeability units, approx. 161, 164, 173,
lithology 8, 264 176, 556
changes of 8, 17 in near-surface rock mass 468
range of- 246 inverse of Qc approximation 226, 555, 556
lithosphere 227, 241 LWD logging while drilling 301, 302, 306, 312–320
oceanic 228 for AVO interpretation 315
thrusting 227 for OBS tie-in to 4C acquired data 315
loading-unloading for horizontal well sections 315
behaviour, intact rock 38, 59 for shear-wave anisotropy analysis 313–315
behaviour, rock masses 97, 98 for warning of pore pressure changes 315
effect on E-modulus 300 velocities, compressional and shear 318
joints 485, 494, 495, 516 velocities compared with wireline 318
hysteresis 182, 300, 514–516
thermal 513–517 Macro
loam 26 deformations 421, 427, 428, 463, 484–486
logarithmic 537, 538 displacements, micro-displacements discussion
logging 492, 493
-based FZI 336 fractures 441
dipole- 312–316 permeability 356
field- 469 pores, inter-particle (oolitic lst.) 348
while drilling, see LWD magma
wireline 302, 306, 313, 318 axial- lense 281
log-spiral magma chamber
failure/fracturing 130, 303, 305, 386 mid-crustal- 281–282
Index 685

magnetometer partly- rock 281


relation to true north 314 meso-scale fractures, Chapman model 465–474
magnitude modelling reservoir anisotropy 472, 474
earthquake- 213–219 role in earthquake studies 470
-frequency plots (intact, fractured) 199 seismic visibility, contra micro-cracks 499
sorted by- (Qseis) 181 Mesosphere 227, 241
Makurat CSFT test (see CSFT) Mesozoic 25
Mamouth Lakes earthquakes, California 212 basement 222
Manitoba 57, 87, 127–131 rocks 142, 153
mantle 243, 244 meta-anorthosite 11, 13, 70
peridotite 257 meta sediments 243
Upper 241, 245 metal discs (model) 470, 480, 481
Upper-mantle velocity histograms 247 metamorphic
wedge 243 rocks 6, 12, 25, 72, 109, 142, 246, 247, 256, 263, 265, 278
major stress 3 rocks, seismic Q 186
marble 57 unweathered 13, 72
artificially fractured columns of- 81 methane ejection 60
micro-fractured, scattering 195 MHF massive hydraulic fracturing 549
marketplace economics 396 mica 243, 509
marl 19, 161 micaceous inter-layers, Bandis model 310, 311
-sandstone 46, 171, 172 mica schist 57
interbedded 212 micro
marine cracks closed by ␴1 36, 199
environment 373 crack system 22
seismic (ubiquitous term) 232–239, 253–290, 334, crack density in fault zone 429
336 crack, elliptical 465, 467
Massif Central, France 159 cracked 5, 21
masking 16 degree of- 182
massive discontinuities, displacements 5, 45, 416–425
crystal structures 509 deformation compliances 5, 416–425, 499
unjointed rock 4, 6 excursions 184, 506
Masua mine 60 flow 190
mated (see joints, fractures) fracture and macro-fracture orientations 441
matrix fracturing
elastic stiffness- 417, 451, 477 imbalances 5
intact 5 seismic event (see also AE) 131
format 374, 477 seismograms 206
porosity (see porosity) 373 valves 193, 194
reservoir- 393 velocity probe 128
silty-sand 12 micro-cracks 31, 127, 128, 184, 195, 196, 201, 407, 418
stiffness- 416, 417 and micro-voids between sand, clay particles 418
maximum APE model, Crampin (see APE)
horizontal stress (see stress) aspect ratio, stiffness, critique 415, 434
temperature (see temperature) aligned 374
entropy method 84 and joints/fractures cause S-wave splitting 408
McMurdo Sound, Antarctic 253, 254 caused by sampling 397
measurement-window 194 discussion/critique of universal role 412–415
mechanical over-closure (see over-closure, OC, of dominant- 195, 196
joints) stiffness compared to fractures 415
medical profession 52 pressure-resistant 226
melt, melted and weathering 225
fraction 256, 257 micron
mantle- 256 sub-interaction 188
parental- 256 microstrain 184
686 Index

mid parallel-plate 172, 279


-Atlantic ridge MAR (see VP and Q data sets) 261–273, visco-elastic, transversely isotropic 351, 353
287–290 velocity-depth 266, 267, 271, 274, 275, 276
-ocean, axis of ridge 282 modelling (see numerical)
-ocean, off-ridge distance 282 excellent- results 362
-ocean ridge 3, 266 forward 251, 253, 391
-ocean ridge, transverse 273 modelling (see numerical, and 2D, 3D, UDEC, FRACOD,
-ocean seismic investigations 179, 261–290 discrete, distinct)
migmatite 139–142 modulus, moduli (see deformation, and Young’s)
weathered 107, 108 of deformation (see deformation) 227, 339
mineral, minerals (see hydrothermal) 22 dry rock- 296
bridging 384, 406, 527 low- damage zone 313
cements deposited in fractures 384 reduced-, in alteration zone 313
cement injection analogue 291–295 Moho (Mohorovicic velocity discontinuity) 229, 241,
composition 256, 257 242, 244, 255
deposition of 287–293 Mohr stress representation 542
effect on VP – age relations 270, 287–293 Mohr stress transformation
filling 11 failure to account for dilation 519
post-kinematic cements 384 non-coaxial stress and strain 519
synkinematic cements 384 Mohr-Coulomb
sealing 270, 276 based continuum modelling 304, 319
mineralization 270, 287–294, 377–380 parameters (c and ␸) 101, 307, 537, 539
bridging (to prevent closure) 406 parameter combination c plus tan ␸ 304, 461
episode 334, 336 parameter combination c then tan ␸ 304, 461
mine, mining 58, 60–62 strength criterion 304, 305
coal 60, 61 stress transformation 552
equipment 53 stress transformation, modification 553
gold 60 theoretical solution 308
hand-mined 128 moisture content 27, 30
potash 123 Mojave
stopes 224 Desert 90
mini-EDZ (see EDZ) east and west regions 260
minifrac (see hydraulic fracturing) gneiss 261
minimum (ubiquitous term) intrusives 260, 261
stress (see stress) moment magnitude 130
mining-induced (seismic) 224 monitoring
Miocene fluid front- 352–353
limestone 11 4D time-lapse 352
marker 15 monoclinic medium, multiple sets 420
Misasa earthquake, Japan 213 monopole acquisition 316
Mississippian sandstone-shale units 371 for radial variation of compressional slowness 316
MIT Massachusets Institute of Technology 318 transmitter-receiver spacings 316
Mjølner (meteor) impact structure, Barents Sea 404 Mongstad oil storage caverns, Norway 147, 160
Mobile Bay, offshore Alabama, USA 320 Monticello reservoir, USA 90
mobility permeabilities 554, 555
high 183, 194 MONT-1 well, USA 90–92
low 183, 194 montmorillonite 27
model joints (tension fractures) 493 monumental study 249
in weak brittle model materials 547 monzonite 21
model-prototype scaling 547 fresh 19
shear strength envelopes, peak, post-peak 547, 548 weathered 19
strength, stress, displacement scaling 547, 548, 552 moraine 11
models (see numerical) mountain, mountains
elasto-plastic 305 Andes 72, 231
Index 687

Rocky Mountains 231 Natih field, Oman 455–458, 469


-side deposits 153 naturally-fractured
-side screes 153 gas reservoir 382
Ural Mountains 231 near-surface (rock masses) 3, 4, 13, 74, 465
mountainous 33 apertures 390
Mount Davis, Hong Kong 76 clay 25
moving source 53 conditions 10, 74
Mratinje da 301,m site, Yugoslavia 49–51, 99, 170 data on stiffness 423
mud extremes (beach sand, ice) 252–254
pressure induced tensile cracking 299, 321 fully-saturated rocks 211
temperature management 321 geotechnical investigations 15
mud filtrate invasion 295, 302, 316–319 investigations 14
accelerated 303, 305, 307–311 layers of sediment 205
based on porosity-permeability conversion 317 low-Q zone, frequency loss in 378
constant permeability with radius assumption 318 material 7
enhanced by mini-EDZ 317 measurements 69
enhanced by log-spiral shear surfaces 317 ocean-floor velocity structures 261–294
enhanced by sheared joints 317 permeability 390
scenario, tunnel analogy 320 permeability tensors consistent with ␴H max direction
speed 307 442, 531
speed highly none-uniform 307 seismic Q 203–205
mud temperature seismic surveys 4
management 321 tunnelling 9
mud-weight 295, 296–298 velocity structure (refer also VP and Qseis data lists) 8, 10,
constant 306 20, 46, 47, 51, 52, 55, 75, 76, 78, 79, 88, 89, 91,
over-balance 299 92, 93, 203
versus depth 296, 298 weathering effects 19
mudstones 7, 8, 9, 26, 79, 94 Neogene rocks 142
inter-bedded 207–209, 234 neural network to infer commercial prospectivity 384
Tertiary 19, 21, 304 neutron log
multi porosity from- 378
-azimuth walk-away 369 Nevada (nuclear) Test Site 251–253
-component 6, 7, 8, 363 Nevada Test Site, Yucca Mountain 542, 543
-channel (ubiquitous term) 254 New Mexico 123
-frequency 207–209,373 Ngendei
offset, multi-azimuth, 3C VSP 410 South Pacific data 275
physics approach 318 velocity model 275
-source multi-receiver 49–51, 54–67 NGI 53, 55, 58, 74, 88, 147, 148, 443, 452, 469, 518
-wave-form acquisition 310 NGI borehole failure study 303–305, 311
multi-variable linear regression Nick Barton & Associates 307
clay content and porosity, VP 332 nine-component, three-dimensional 9C/3D 455
melt fraction, VP 256, 257 Nirex (see UK Nirex Ltd)
mineral compositions, VP 256, 257 NMO normal moveout 389
multiple ellepticity 390
borehole logging tools 379 stretch 389
fracture directions at faults 394 nomograms (Q, M, K-L, c) 176
position borehole extensometers MPBX 301, 367 non-
MWD measurement while drilling 313 aligned with ␴H max direction, polarization 384, 385,
mylonite 13, 70, 100 389, 390, 403, 404, 406, 414, 429, 431, 432
aligned with ␴H max direction, conductive joints 527,
N and S components 484, 485 541, 542
NaCl (see brine) 342 conductive 531, 541, 542
Nagra 38 linear shear strength criteria 339, 537–540
Nathpa Jakri hydroelectric project, N. India 302 uniqueness 271, 369
688 Index

Norfolk, England 29 Northern Appalachians 231


normal Norway 13, 16, 43, 53, 55, 56, 70, 74, 88, 144, 147, 160,
closure (N) 484, 485 161, 170, 212, 216, 217, 408, 518
closure is least productive condition 488 Norwegian Geotechnical Institute, Oslo (see NGI)
compaction trend 297 Norwegian Petroleum Directorate 401, 452
compliance ZN, dynamic, bedding 500 Norwegian Road Authority 147
fourth-cycle loading concept 484, 487 nuclear waste
loading of joints 484, 485, 487 disposal investigations 124–136, 161, 363
loading and hysteresis 484, 485, 487 related research 165
normal stress 79, 198, 199 numerical models, modelling
closure BB model 487 BB modelling of coupled stress-closure, shear-dilation, aper-
closure, shear/dilation, aperture e, permeability, BB model- tures 531–533
ling 531–533 BEM boundary element method 306
deformation, rock plus joint 495 Cellular automaton model, scattering, Vlastos-Narteau
permeability BB model 487 463–465
normal stiffness 17, 198, 199, 202, 485, 487, 494–499, Cundall continuously yielding joint model 460
515–517 Cundall distinct element modelling developments
and shear stiffness Kn, Ks 485, 492–499 460, 461
and shear stiffness, clay-filled discontinuities 505 distinct element DEM (finite difference)- 306, 308–310,
apparent reduction with temperature 516, 517 421
apparent ‘zero’ with thermal OC 516 elastic flexural 257.
-compliance discussion 421–425, 492 elasto-plastic flexural 257
constant in DST 490, 492 FEM, 2D, with Goodman joint elements 460
dynamic versus frequency 498 FEM, 3D, dam foundation modelling 460
dynamic/static data, stress-dependent 496 finite element, jointed- 421
estimated from block test 515–517 FRACMAN, Dershowitz, Golders, 3D fracture-flow
fracture zone in URL 506 code 460
high stress levels 495, 496 FRACOD 306, 307, 461
initial (Kni) 494 geomechanics 1D-strain model, Ekofisk 399, 400
interlocked joint 494 HMT fully-coupled, water-flooding in fractured medium
mismatched joints 494 402
normal stress 498 NAPSAC, AEA Harwell, 3D fracture-flow code 460
reduction, apparent, with temperature 515–517 orthogonal sugar-cube model, permeability-seismic
-shear stiffness ratio, dynamic 497 integration, Brown et al. 479, 480
-shear stiffness ratio, static, scale dependent 498, poro-elastic (see next)
499, 502 synthetic jointed/fractured reservoir models 393
normal moveout (see NMO) 3DEC 365, 454, 460
normalized surface of invasion 360 UDEC (see UDEC) 365
normally-pressured 352 UDEC dynamic attenuation 478, 479
Norris Lake Community, Georgia, USA 215 UDEC-MC 306, 308
North America, American 231, 246 UDEC-BB 460, 469
deep gas reservoirs 320 UDEC-BB, HM block test 518, 519
plate boundary 258 UDEC-BB, deformation of jointed rock masses 484–486
reservoir rocks, stress magnitudes 298, 299 UDEC-BB, reservoir compaction modelling 401, 420,
North Anatolian Fault Zone, Turkey 220 443, 452, 453
North Cape Tunnel 147 UDEC-MC, reservoir subsidence modelling 400, 427, 428,
North Caucasus oil fields 450–452 454
North Sea 352 UDEC-BB, tunnel/borehole modelling 306, 309, 310
reservoirs 235, 311, 312, 391, 396, 397, 398, 412, 420, numerical poro-elastic models 236, 377, 391, 407,
427, 447, 451, 452, 504 459–477
salt dome (Zechstein) 381 Angerer, shear-wave anisotropy changes 451
sands 326 Biot squirt-flow attenuation 236
sandstone 359 BISQ Biot and squirt-flow model 378
UK-sector 235 BOSK effective medium model 457, 458
Index 689

Chapman, triple-porosity 405, 415, 465–475 prices per barrel 320


Hudson effective medium model 461 sand 330
Nishiwaza TI cracked medium model 362 saturation changes in 4D 397
SeisRox visco-elastic model, Johansen et al. 375, 376 saturation mapping in 4D 397, 398
super-k poro-elastic- 377 storage cavern 53, 54
Tod effective medium, crack decay model 462 -from gas, distinguishing 455
Nurec deep borehole data 555 -to-gas conversion 296
well (see wellbore, borehole)
OBC ocean bottom cable 391, 394, 396, 397, 452 oil sand
hydrophone 391 heavy-, steam injection time-lapse 379
geophone, in-line, vertical, cross-line 391 effect of temperature on VP 330
OBS ocean bottom seismometer 237, 281 effect of oil/gas %, Kern River- 330
observed rock effect of oil/brine %, Venezuelan- 330
classified not characterized 310 oil field, fields
ocean complexity of recently discovered- 323
-bottom hydrophone OBH 275, 281 heterogeneous distributions of parameters 323
(analogue AOBH, digital DOBH) 281 spatial variability 323
-bottom receiver array 281 spatial variability of porosity, clay-content, fracture density
-bottom seismic instruments 254, 261 323
depths 227 oil-water contact 369
Drilling Program ODP 206, 276 Oishiyama earthquakes, Japan 210, 211
floor (ubiquitous term) 261–293 Okayama, Japan 48
floor rock quality 285 Oklahoma, SE 374
sub- basalts 261–293 oldest units 369–371
oceanic crust olivine crystals
fracture zones 271–272 aligned- in Upper Mantle 276
young 276, 282 Olkaria geothermal field, Kenya Rift 394
oceanic lithosphere 228 Oman 455–458, 469
age of- 228, 262–265, 287–294 one-dimensional strain compaction modelling 399–401
down-hole sonic logging 261 of unjointed chalk 400, 401
old 228, 287–291 onshore sites 19
young 228, 287–291 ooze-chalk transition 206
oceanic sub-ocean layers open joints parallel to ␴H max assumption
Layer 1, 2, 3 243, 244 (see fractures-open, parallel, stress ␴H max, non-aligned)
Layer 2A, 2B, 2C 277, 283–285 options for open joints, fractures 549
Layer 4, 5, 6, etc. sub-ocean divisions 244 ophiolite, on-land, Troodos, Cyprus 278
Oceanografer, North Atlantic fracture zone profile 273 O/R open, rock-to-rock contrary rotations 414, 446, 450,
Oddatjørn dam site, Norway 170 453, 527
offset possible source of polarization rotation 414, 446, 450,
and azimuth, variation of seismic data 382–396, 474, 525–527, 550, 551
401–406, 407–482 (Ch15) Ordovician sandstone-shale units 370–372, 370
by shear mechanism 414, 518, 523, 524, 541–543, ore bodies 22
550, 552 organic matter 352
fully-populated- 373 oriented
large 273 core (see core)
small 273 orientation
offshore discrepancy re anisotropy axis and stress 383, 384
geophysics 94, 243–245, 251, 253–294 discrepancy re reflection S1 and polarized S1 475
regions 260 unfavourable 14
oil orogens (geologic structures) 246
bearing rock (ubiquitous term) 233 orthogonal (joint/fracture) directions (three) 25, 172,
bearing sandstone reservoirs 343 353–363, 479, 480
dead- 296 orthorhombic material 374
live- 296 orthotropic 46
690 Index

oscillating 5 phase velocities 480


point (sources) 153 reduction of P-wave velocity due to rock failure 36
Oslo signal propagation 481
downtown 139 survey, surveys 70, Ch14, 384, 526
Fjellinjen Tunnel 54, 139 travel time, temporal 433
numerical model 469 travel-times, for AVO 388
fjord 56, 217 travelling obliquely to fractures, shear properties of rock
Otsuki fault zone, Japan 47, 48 404
outcrop mapping 76, 455, 456 velocity anisotropy 38
output signal 118 velocity contrasts 8
overburden (see also stress) velocities 11
layers 15 Pg and Pn direct and refracted waves 242, 246
gradient 296–300 P-S converted (C) waves 369, 388–390, 438–440, 452
stress 299 polarity reversal across strike 440
stretching of-, subsidence 399, 400 pac-ex packer-extensometer unit 505
velocity anomalies 14 Pacific plate 243, 245, 258
over-closure, over-consolidation boundary (southern California) 258
episode, in DST preparation 517 Pacific ocean 243, 245, 246
heated in situ block tests 513–517 sub-ocean East Pacific rise 281–290
heated lab tests 513 western 206
mechanical- (of joints) 87, 513, 517 packer, packers
ratio, effect on shear strength 517 double- 162
thermal- (of joints) 87, 513–517 quadruple- 174, 292
over- pressure, over-pressured (see pore pressure) 295–298, test 378
350, 351, 352, 450, 452 Pahute Mesa, Nevada Test Site 251–253
detecting-, factors 350 Paleocene 447
example from tunnel 319, 320 Palaeogene rocks 142
reduced velocity-depth gradient due to- 450, 452 Paleozoic 25, 72
thinly-bedded strata with- 320, 452 rocks 142, 211
time-sag due to- 381 sediments, Ko 299
top of- 297 parallel
shale, lab tests 354–356 bedding- Q 352
zones 268 not- to stress direction 384
Oxford clay 49, 51 to aligned fractures, faster P-waves Ch3, 382
Oxfordshire 77 to bedding 354–357
oxides to fractures, minimum influence on reflection 388
SiO2, MgO 256, 257 to major jointing 35, 46
FeO, CaO, Al2O3, Na2O 256 to schistocity 38, 39
to ␴H max assumption 179
P-wave (see VP annotated results) 4, 5 to ␴H max assumption (microcracks)
amplitude 81–86 to H max assumption (open joints, fractures) 384,
amplitude/magnitude spectra 198, 199 526, 541
and S-waves 6, 7 to stress direction 35
and S-wave anisotropy 37 parallel and/or perpendicular loading (and ray paths)
and S-wave surveys 70 w.r.t. layering, bedding, schistocity, foliation 38, 39, 46,
and S-wave velocities 6 354–357
anisotropy (azimuthal) 35, 179, 382–406, 438 w.r.t. extension fracture 362
anisotropy definition 438 w.r.t. fracrure orientation 389
axial- 36 seismic Q in fractured shale reservoir 236, 237
cancelled between fractures 466 parameter
energy 130 combinations 271
joint use of- and C-wave 396 non-unique- combinations 271, 272
multi-directional 38, 39 ratings (see Q-value, see various J-parameters, see Appendix A)
particle motion 5, 438 Paraná Basin 231
Index 691

Parkfield Dense Seismograph Array, (HRSN) USGS, depth dependent- 164, 490, 555, 557
California 216, 223, 224, 429, 431, 432 dominant direction, fracturing 455
particle enhanced 133
motions 5, 437 e2/12-based- 173, 474, 513
size (grouting) 487 enhancement with pre-peak shear 554
velocity histories 479 estimation from QH2O 554–558
passive source (see AE acoustic emission) FZI data 333–335
patchy shooting 391, 394 high-permeability zones 17
patchy saturation 335–337 influence of conjugate shearing on- 358
causing frequency dependence 336, 337 influence of shear stress on- 517–536
capillary pressure-saturation estimates 336 low-, due to shale 236, 237, 296
discussion re compliance, squirt, attenuation 336 lower 151
due to rock joint 336 Lugeon test of- (see Lugeon)
due to macroscopic patch 336 Lugeon – K conversion, approx. 161, 164, 173,
mixture 336, 337 176, 556
with ‘homogeneous’ mix of lithologies 336, 338 macro- 356
path length 30 maintenance of joint- 358, 359
Pb3O4 red lead, model additive 547, 548, 551, 552 maintenance due to shear 402
PC-element matrix- 373
liner 151 matrix-, normalized 360
ring building (TBM tunnel) 94 max. and min. 174
peak shear strength (see joint, shear strength) measurement facility 359–361, 363, 364
singularity of ␦peak and beyond peak 554 measurement value, if intersecting structure 407
peat 26 micro- 202
pebbles 12 of matrix too low, Ekofisk 401
pegmatite 13, 70, 73 of mineralized ophiolite (estimated) 278
weathered 107, 108 of rock mass 173–175, 292, 293
Peko Oil USA 374 of rock mass, one set 533
penetration orthogonal- 358, 359
rate 78, 139, 149, 158, 321 parallel to joints, fractures 358, 359, 362, 363, 363
strength 79 principal magnitudes (tensors) 278, 291, 292
Pensylvanian sandstone-shale units 370–372 Qwater, modification for- 556
performance of TBM 158 QH2O – depth – permeability 557, 558
peridotite 19 ratio of principal-, in situ 174, 292
permeable 169 ratio of principal-, around borehole 318
fractures 179 sandstones compared 343
sub-ocean crust 266 scatter curves 334, 335
zones, Qseis/VP correlation for sands 376, 377 stress behaviour 179
permeability, permeabilities 77, 159–165, 169, 170, 203, stress behaviour (in situ tests) 165, 166, 527
331–337 stress behaviour (lab tests on coal) 168
and storage 364 stress behaviour (large-scale lab) 527
anisotropic- 358, 359, 382, 425 super-k method of prediction 378
anisotropic-, Äspö spiral tunnel 555, 556 tensors 359, 360
access spiral tunnel tensor rotation 174, 292
as function of confining pressure 359 test, testing 9, 88, 159–176, 292
as function of direction 359 test holes 513, 514
before and after grouting 173–175, 292, 293 three-axis 357–364
changes due to tunnelling 615 variable- (crustal scale) 231
-clay-content 331–337 -velocity behaviour 168
comparisons 278 virgin- 278
core-plug- 237 Permian sandstone-shale units 371
depth dependent perpendicular and parallel loading (and ray paths)
deep well- data, land-based 555 w.r.t. layering, bedding, schistocity, foliation 38, 39, 46,
depth data (1600 m of oceanic crust) 279 354–359
692 Index

Perspex (see Lucite) pillow (see also basalt)


buffer rod 501, 502 lavas 277, 279
perturbation of stresses, properties in EDZ 302–321, pilot drilling 150–152, 319, 320
615 pipeline 3
Petatlan earthquake 213 Piper Alpha platform disaster, UK sector of North Sea
petite-sismique 110, 112 320
petroleum Pirapora dam site, Brazil 173, 174
exploration for deeper – reserves 320 plagioclase 243
geologist 369 plastic
industry 301, 323, 523 deformation 93, 303
industry benefit from shearing mechanism 549 failure (irreversible) 183, 303
oil prices 320 model material 302
source rock 351–353 replicas of shear-dilation path 527
well surveys 369 zone 140, 303
petroleum reservoir, reservoirs (see VP and Q data) 232–239, plate
369–406, 438–459, 471–475 (jacking) load test 46, 50, 51, 74, 97–105,
conventional wisdom 531 112–114, 122
depth-log of Q -VP -K fractured shale 237 parallel- (see aperture)
familiar warmth of a-, re thermal OC 516 tectonics 179, 226, 227
‘open’ joints, difficulties with conventional direction plate tectonics structures
526–536 Benioff zone 227
‘open’ joints from mineral-bridging 384, 406, 523 bulge 227
‘open’ joints from shear and dlation 523–536 forearc basin 227
plot of depth-1/Q-VP 233, 234 remnant arc 227
plot of low Q, high Q1 in fault zone 235 seismic belt 227
plot of Q-VP 234 subduction complex 227
plot of Q in anticlinal chalk 235 trench 227
pressure-depth-gradient-buoyancy aspects 296–297 volcanic arc 227
rocks, rough, hard end of spectrum 515 platforms
where rock strengths are limited 384 geologic structures 246
petrophysicist 369 jack-up of- due to subsidence 400
phase velocity of wave 198 Pleistocene rocks 142
phi-r r residual friction (see friction) points of contact (joints) 30
Phillips Petroleum Company (Conoco Phillips) 452 Poisson expansion 133
phyllite 12, 32, 38, 247 Poisson’s ratio (dynamic, from VP and VS) 5, 6, 7, 8, 9, 57,
quartzitic 122 100, 104, 106, 107, 109–111, 130, 161, 263, 264,
slatey 122 270, 280, 337–339, 363, 394, 396, 529
unweathered 12 azimuthally-dependent-, 0 to 5 km, Kenya Rift 394,
weathered 12 396
physicists 184, 282 anomalous 270, 337
physical -depth, sub-ocean Hole 504B, Costa Rica ridge 280
aperture (see aperture) depth, 0 to 5 km, Kenya Rift 396
laws of behaviour 16 differential pressure 338, 339
models (borehole drilling) 303, 305 dynamic (function of axial stress) 126
model (poro-elastic, dual porosity) 470, 480, 481 effect of low differential pressure on- 339
models using tension fractures 523, 524, 551, 552 effect of hydraulic fracturing on- 338, 339
Piani di Ruschio dam site, Italy 101 effect of brine, crude oil, gas/dry on- 338, 339
Picrocol data set at salt dome 380, 381 extremely high- , below ocean floor 286
piezoceramic vibrator source 61 lateral expansion, rock masses 485
piezoelectric mass-, lateral expansion exceeds 0.5 523, 524
bender transducers, high resolution 375 pseudo-static 9, 105
source 52 theoretical relationships 6, 7, 104, 338
transducers 192, 292, 361 tomogram 379
vibrator 62 Poisson’s ratio – depth (shallow) 8, 111
Index 693

Poland 60 porosity 9, 13, 27, 29, 270


polar age-depth, young oceanic crust 263, 276, 369
diagram, 1/Q of fractured medium 405 bi-modal- 349, 350
histogram 385 correction 92, 93, 163, 167
polarization (see shear wave) critical, suspension limit 324
direction not parallel ␴H max 432 dual- 369
fault-parallel alignment 432 dual- physical model 480, 481
inversion, orientation 404 equant 465, 467
90°-flips of- 415, 416 high- sediments 206
90°-flips of-, alternative, axial over-load 416 hard 272, 276
parallel to fast formation axis 314 lack of crack- 31
rotation due to deep well injection 413, 414 loss of- at extreme depth 311, 312
rotation due to O/R concept (see O/R) 548–551 matrix 24, 285, 373
shear-wave arrivals 448 ranges of- 311
temporal changes in Cornwall HDR 413–415 -reduction due to water saturation weakening 399
polarized shear waves (see shear waves) sandstones compared 343
polyaxial stress state loading 302–305, 358–364, 419 secondary, crack- 270
polyurethane foam 305 soft 272, 276
Pont Ventoux total 21, 22, 24
hydroelectric project, Italy 154 porosity – uniaxial compressive strength (crystalline and
headrace tunnel (TBM) 154–157 volcanic rocks) 28
polluted sludge 25 porosity – uniaxial compressive strength (diverse soil,
Ponte Cola dam site, Italy 101 rock) 26
pore, pores porous
collapse, accelerated due to water 400, 402, 442, 540 granular media 23
collapse and fracture stiffness discussion 426, 538, 540 macro- 373
compressibility 452 micro- 373
filling minerals/materials (clays) 331–337, 349 rock (ubiquitous term)
flat 190, 205 porphyry 13, 70
fluid 5 strongly jointed 148
geometry 356 porphyrite 20
size distribution 332 Portugal 170
space compaction 296 post-peak region of stress-deformation curve 36
space compressibility 351 post-stack seismic time section 14, 15
space compromised by depth 295 potash 123, 124
space occupied by ice 33 power law, frequency components 205
water flow, attenuation 282 Pratt-Swolfs in situ block tests, USA 165–166
pore pressure 17, 169, 295–298 precipitation (see sealing, minerals)
analyst 295 pre- and post-fracturing
change, prediction of away from wells 397 effects on VP and VS components 362
coefficient 268 effects on 1000/Q 363
-decline, effect on bulk modulus 397 effects on permeability 364
divided by confining pressure 355 pre- and post-peak displacement condition
effects 268, 355, 356 permeability enhancement 554
elliptic propagation 464 relation to resisted mobilized friction magnitude 554
excess- 298, 352 pre-Cambrian 211
excess-, effect on QP 352 basement 429, 430
excess- effect on VP and VS 298 crystalline, Ko 299
gradient 296, 297 preferred orientation
independent application of- 355, 359 of clay particles 354, 355
reduction prior to water-flooding 401 pre-injection/pre-grouting (of cements) 4, 56, 118, 153, 172,
regime 273 173, 175
poroelastic modelling (see numerical) analogue for hydrothermal sealing, VP – age relations
poroelasticity 209 287–291, 291–294
694 Index

pre-injection/pre-grouting (of cements) (contd ) Q-logging of outcrops 94


pre-peak region of stress-deformation curve 36 Q-logging of tunnels 94
pre-slot (unloading of block) 167 Q-parameter ratings re rock quality, (see J-parameters, see
pressure, pressured 30 Appendix A) 615–624
chamber test 97, 107, 108, 115 Q-prime (Q ) 469
dependence (ubiquitous term) 346, 348 Q-jumping (quality-depth; see curve-jumping) 93, 268,
excess- 296 272, 275, 285, 286, 625
lack of pressure sensitivity 31 Q-system of rock mass characterization 4, 250,
over- 296–299, 350, 351 Appendix A
pore- (see pore) 296, 297 Q-system for selecting support 163
-tunnel 119 Q-system support pressure 140
under- 351 Q-variation with depth 386, 387
primary joints 551, 552 Qo oriented rock quality 411, 615, 616
principal
directions of loading 355, 357, 359, 361–363 Q-values (logged or estimated) 6, 7, 11, 13, 16, 17, 18, 43,
directions of velocity measurement 47, 48, 53, 56, 58, 69, 70, 74, 75, 76, 78, 88, 92, 93,
stress (see stress) 361–363 113–115, 131, 132, 141, 144, 147, 148, 150, 153,
prismatic sample, 16-sided 481 157, 158, 160–164, 167–172, 203, 225, 246, 250,
probe drilling 139, 151–153, 162 268, 273, 285, 287, 348, 377, 387, 425, 447–450,
processing software 7 469, 499, 500, 515, 556–558, 615–626
production assistance 438 Q-values converted to Qc (normalized by UCS) 92–95,
production rates 114, 144, 148, 209, 215, 216, 220, 225, 246, 250,
damaged by excessive- 363, 453 252, 259, 268, 272, 274, 285, 287, 348, 625
effect on jointing 487 Qc – M 92, 108, 113, 114, 132, 134, 269, 425, 449,
profile (see roughness) 625, 626
propagation Q-VP–UCS 449, 625, 626
direction 353–363, 451 Qc-VP relationship 6, 13, 18, 47, 74, 92, 93, 108, 134,
increase of velocity with non-vertical- 451 147, 148, 478
of body waves 5 Qc-VP –depth 93, 157, 233, 271, 290, 447, 449, 625
of stress waves 17 Qc-VP –depth – porosity model 92, 134, 157, 233, 268,
proton accelerator foundation 29 450, 452, 465
Prodozakonov f-value 76 Qc-VP –Lugeon relationship 159–165, 167, 170, 171,
psuedo-static 172, 175–177
macro-deformations 6, 421–425, 456 Qc-VP –M (depth) model 161, 167, 232, 257, 258, 348,
normal and shear stiffness in CO2 flood 451 387, 449, 499, 500, 515
parallels 9 Qc-VP –M–L model 163, 164, 176, 425, 556
pulse echo method 346 Q-VP–PR trends (TBM tunnelling) 151
pulse generator testing 350 Qo 615, 616
pumping test (extraction) 162, 173, 174, 378 Qtbm rock-machine quality factor 17
Pure and Applied Geophysics 230 Qtbm model 149
push-down 14, 15 Qwater, QH2O – depth model 556–558
push-up 14, 15
pyrite 23 Q seismic quality (also refer to attenuation, earthquakes)
pyrrhotite 23 (see itemized data below alphabetic order)
Qseis (and attenuation 1/Qseis ) analysis methods 238, 239
Q rock mass quality (termed’ rock quality Q’ or Qrock ) Azimi second, third laws 239
179, 269 Cole-Cole 239
Q-calculation example 225 Futterman causal operator 238
Q-calculation example, Sugar Loaf, Rio de Janeiro 558 Kjartansson 239
Q-calculation example, faulted rock, tunnel collapse 558 Kolsky-Futterman 239
Q-histogram core logging 224, 622–624 peak amplitude ratio 238
Q-logging, absence of 208, 224, 386 power law 239
Q-logging of caverns 53, 94, 625 pulse-broadening technique 238
Q-logging of core 74, 94, 274, 386, 387, 623, 624 spectral (amplitude) ratio method 199, 201, 238, 386, 464
Index 695

rise-time method 238 QS (earthquake-based)


velocity dispersion formula 386 QS – depth (Cajon Pass, Varion wells, Parkfield) 222–224
Qseis (Q or QP or QS) (alphabetic order) 17, 66, 179, QS – depth (0–35 km) and lateral variation
181–239, 342–353, 354–357, 363–367 (0–200 km) 229
anisotropy near-surface 300 QS total – frequency (earthquakes), 223
anomaly, anomalous 204, 205, 342 Qscatter – Qintrinsic (earthquakes), 213, 214, 224
crustal lateral variation (also QS) 229 QS fast, QS slow 0–200 m depth, Chi-Chi 436
elastic components of attenuation1 Qe and Qk and Qp
Mid-ocean ridge QP, QS
and Qs 193, 194
QP – QS – depth East Pacific Rise 14°S 286
enhancement by tangential stress 301
QS – depth (hole 504B, Costa Rica ridge, 0 to 1600 m,
extremely low Q 179, 205, 227
also Qintrinsic) 280
high-Q 179, 227
1/Q – depth (sub-ocean basalts, mid-Atlantic ridge, 0
laboratory samples, sandstones 186–195, 341–351
to 700 m) 269
laboratory joint/fracture samples 200, 361, 363, 364
1/Q – depth – laterally (sub-ocean basalts and magma cham-
low-Q 179, 227
ber, 3D tomography) 282
near-surface gradient 211
near-surface- (rock) 201–203 Reservoir Q
Qgas/Qbrine versus frequency, BOSK model 458 Q – depth (sedimentary rocks to 2.4 km depth) 232
Qseis from VP , finely interbedded rocks 380 Q – interval stacking velocity (VP) sedimentary reservoir rocks
Qseis from VP from Qrock 377 234, 235
Qseis proportional to Edyn 387 Q – VP equation: fit to some sedimentary rocks 235
Qseis/VP ratios, distinguish sand, jointed sst. 376 Q-VP – K log, fractured reservoir, shale, 4000–4350
sub-ocean sediments (ooze-chalk) 206 feet 237
unconsolidated sediments 205, 206 QP and QS ranges, sandstone reservoirs bearing: gas, gas con-
variation, erratic, of QP and QS with depth, as Q, Emass 386 densate, oil, water 343, 345
variation with confining pressure 192 QP – QS – porosity differentiation, gas zone indicator in well
variation with frequency 200 345
Qseis similarity of magnitude to QP/QS crossed behaviour with gas contra water/oil saturations
static Eintact (if GPa units) 191, 192, 342 343, 349
static Ejointed (if GPa units) 200, 201, 362 1/Q – depth, anomalously low (Lower Triassic sandstone at
static Emass (if GPa units) 202, 203, 210, 220, 221, 4 km depth) 233
224–226, 232, 269, 348, 350, 352, 365–367, 1/Q – frequency, modelled reservoir fractures 238
377, 379, 387, 405, 410, 411, 499, 500 Q⫺1 in North Sea fault zone, anomaly 235
Non-earthquake field results for seismic QP and QS
Earthquake sources
QP – freeze-thaw cycles, Jurassic limestone, relevance
Qc coda (earthquake sources) 209–219
to field 185
Lg coda 230, 231
QP – depth into tunnel wall (EDZ), columnar basalts, 40 m
Qc coda – frequency 217
depth 127, 367
Qc coda – temporal variation 215
QP – angular frequency and shear strain level, clay 189
Qc coda – before/after earthquake 218
QP – frequency and geophone interval variation, sediments
Qc coda-azimuth-depth-distance-magnitude-time 216
205
Qc1 coda attenuation related to coda wave 209–219
QP – QS dual-porosity chalk, low values near-surface 350
Qc1(earthquake sources)
1/Q (1/QP) – frequency data comparison using Biot 209
Qc1coda – frequency data sets 211, 214, 218, 228
Qc1coda – temporal variation 218 Petroleum related field results
Qc1 coda – before/after earthquake 218 QP – depth slice (mean 32, 96, 106 m and laterally, quater-
QS versus QP coda components 212 nary, alluvial stream beds, lithified sands-sandstone)
Qo (Lg coda at 1 Hz) 230, 231 374, 375
QP (earthquake-based) QP – pressure (equivalent to depth) ultrasonics (interbedded
QP – depth (Cajon Pass, Varion wells, Parkfield) 222–224 sand-, lime-, silt- and mudstones) 207
QP – depth (0 to 35 km) and lateral variation (0–200 km) 229 QP – depth w.r.t. sonic logging (interbedded sand-, lime-, silt-
QP – depth (0 to 1200 km) and VP, from a) body waves, b) and mudstones) 207
surface waves QP – depth, also VP – depth, also QP /VP – depth, 210–400 m,
QP – distance – magnitude (NAFZ) 220 sands and sand channels 375–377
696 Index

1/Q – depth (sedimentary rocks to 2.4 km depth) 232 1/QP – frequency, calculated for fractures/joints of different
1/Q – depth (0 to 5.5 km depth, Lower Triassic sandstones stiffness, quartz-monzonite 200
deepest 1.5 km) 233 1/QP – permeability: various limestones: packstone etc. 349
1000/QP – frequency dependence(interbedded sand-, 1/QS – permeability: various limestones 349
lime-, silt- and mudstones), field 207
Rock physics/poro-elastic modelling
Rock physics data (sandstones) 1/Q – frequency, fracture nucleation, automaton model 463
QP , VP – frequency – effective pressure, dry or brine saturated, 1/Q – frequency,Chapman, fracture size variation 471
Berea sst. 344 1/Q(components) – frequency, SeisRox TIH model 476
QP – differential pressure, Berea sst. 352 qS1 and qS2 examples parallel and perpendicular 390, 445
QP , VP – frequency – effective pressure, dry or brine saturated,
quarry 27, 28, 42
Boise sst. 344
blasting (damage) 170
QP,, VP – frequency – effective pressure, dry or brine saturated,
quartz 27, 243, 254, 509
Massilon sst. 344
cement, precipitation of 311
1000/Q and Q – 3D confining pressure – permeability, cubic
content in three sandstones 343
specimens, Penrith sst. 363
diorite 27
QP – clay content (volumetric) – permeability, sandstones,
mineral fillings 278
equation 346
monzonite (Stripa Mine) 134–136, 422–424
QP – differential pressure, as additional function of frequency,
monzonite (Stripa), joint behaviour 197–200
Berea sst. 187
monzonite, stiffness, dynamic/static data, stress-dependent
QP – % compliant minerals, effective pressure 5 or 60 MPa,
496
sandstones, siltstones 347
quartzite, quartzites 13, 38
QP and VP – effective pressure, microcracked sandstone
bedded, steeply 301, 367
347
jointed 93
QP and QS – differential pressure, Berea sst. 342
Ko 299
QP/QS – (VP/VS)2 and pore pressure, to distinguish sand and
massive (South Africa) 224
sandstone 351
quartzitic
QE – confining pressure and saturation, resonant bar
sandstones 144
technique 186
quaternary
QE – saturation and frequency, resonant bar technique 190
active faults (Japan) 217
QE – strain amplitude, resonant bar technique 187
deposits 447
Rock physics data (various rock types)
QP – axial stress (sat. and dry), basalt 127 R equivalent roughness of rockfill 538
QP – pressure: sandstones 193, bedded-coal 201 Rabcewicz 76
QP – pressure, also Ppore/Pconfining 0°, 30° and 90°, radial
over-pressured shales 355 rock property variations 316
QP – porosity: igneous and metamorphic rocks, limestones and stress (see stress)
sandstones 186 unequal- stress gradients 315
Qph and Qpv – confining and pore pressure, sandstones rail tunnel, high-speed 17
192 Rangely anticline, Colorado 167
QP – compliant minerals %: 5 and 60 MPa confinement, ray
siltstones and sandstones 347 curved- paths 44
QS/QP – degree of saturation, lab 189, 232 paths 16, 59, 61, 62, 130, 261
QS/QP – VP/VS – degree of saturation, sandstone 189 paths parallel and perpendicular to assumed fractures
QP – QS calcareous-, dolomitic-, siliceous-limestones 349 384
QP – QS – oil or water saturated: limestones, bi-modal paths crossing max. no. joints 124–127, 367
porosity 350 tracing 153
QP – QS – kerogen %: modelled Kimmeridge shale, 0° and Rayleigh
90° 353 scattering attenuation 196, 480
QS – axial stress (sat. and dry) basalt 127 wave scattering 480
1/Q (1/QP) – axial stress-strain, loaded to failure, tuffaceous rebound hammer (see Schmidt) 101
sst. 195, 196 receiver, receivers
1/Q – pressure (with three Q⫺1 anisotropy parameters 357 lines (3D-4C) 391
Index 697

permanently installed in seabed 396 reservoir


reconstructed shear-dilation path 527 anisotropy investigations 404–406
recording characterization 295
station (earthquakes) 210–225, 261 characterization using dispersion 387, 388
vessel/boat 396 compaction 183, 295
recovery factor 373 completion 295
recrystallization of firn 254 depletion phenomena 397
reflected detection of- compaction in 4D seismic 397
P-wave 438 detection of- subsidence 4D seismic
primary 383 drawdown (dam sites) 88–90
S-wave 438 description 295
upgoing multiple 383 dynamic- properties 397
reflection 283 engineer 369
amplitude changes in 4D 352 fractured/unfractured water-flood survey 402, 403
amplitude varies with azimuth AVOA 389 heterogeneities 372, 373
amplitude varies with offset AVO 389 horizons 297
coefficients, azimuthal P-wave 393 impounding (dam sites) 88–90
coefficients, for split shear waves 393 management 394, 396
coefficients in one- and two-set models 393 parameters, estimates of from 4D4C 396
coefficients in AVO 388, 389 phenomena 179
enhanced-, subtraction of seismograms 382 pressure 295–299
methods 4 production 295
seismic, comparison to polarisation 475 production increment 396
seismic, exotic uses 403 rock scenarios 535
-strength change, in 4D 398 residual
wide-aperture seismic- 285, 286 friction angle (see friction) 507–512
reflector 249, 255, 383 shear strength (see joints, fractures) 507, 522
deep-, difficulty of imaging, sub-basalt 403 resistivity
in-tunnel 153 – depth plot, sub-ocean Hole 504B, Costa Rica
refraction seismic ridge 280
inversion 16 logging 169
land-based 274 resistivity – depth relation 20, 297, 298
on-bottom- 282 temperature effect 297
reversed deep-sea (historic) 242, 243 tomograms 169
shallow 3, 4, 5, 9–16, 74, 160, 203 Reskajeage, Iceland 503, 504
refractor 4 resolution
imaging 15 -problem 7
refuse transfer cavern 76 vertical 7
regional variations of Q 231 resolving kernels 262
regression lines 171 resonance
relaxation 182 decay measurements 187
mechanisms 183 extensional- tests 186–188
relaxed 183 resonant bar techniques 187–190
-pore fluid in super-k regime 378 REV effect 223
reloading of fractured cube 363 rhyolite
repeated surveys (see 4D) artificial surfaces 82
replicas of shear-dilation paths 527 ridge (see mid-ocean)
remnant arc 227 rift structure
Rendalen hydro electric project, Norway 147 Rio Grande- SW USA 230
research, experimental (ubiquitous terms) earthquakes, New Mexico 210
borehole 207, 303, 305, 308, 346, 356 ripping 30
reserves river 162
efficient exploitation of- 396 rms (root mean square error) 472, 474
698 Index

RMR -rock mass rating 7, 11, 56, 70, 76, 144, 163, 469, stresses, principal components 320
615, 621 support, high crack density 444
core-logging 387 type 9, 11, 27, 173
correlation to static deformation modulus (Ed or M) types for JRC, JCS, ␸r determination (11) 509
111–113 unstable- 160
correlation (approx.) to rock mass quality Q-value 113, rock engineering 3, 30, 106, 215
163, 621 parallels 226
-variation with depth 386, 387 parallels to sub-ocean gradients 286
rock project, projects 35, 110, 423
anchor foundation 51, 52 rock failure 127
bedded 304 dilatant- 302–312
beneath our feet 558 non-dilatant- 302–305
blocks 156 rock mass 30
boundaries 202, 226 characterized not classified, outside EDZ 310
burst prone areas 52 characterization method 202
caverns, models 551 classification method, Chinese 74, 76
classes 139–141, 152 condition 9
classes and rock types 142–144, 148–150 deformability (pseudo-static) 97–115, 161, 162
conditions (adverse for tunnelling) 151 failure (under plate load) 102–104
country- 157 near-surface-, attenuation 226
cover 147 parallels (shallow crustal seismic attenuation) 224–226
deformation, large scale 400 quality 6, 7, 17, 92, 95, 225, 273, 282
fabric orientation, from Point Load 384, 385 quality improvement (due to grouting) 170–175,
finely-layered sequence of 207–207 292, 293
framework 356 quality, low 158
hard jointed 164 rapidly changing- qualities 10, 203
hard massive 93, 164 rapidly changing- qualities at fault zones 394
hard massive, completely intact, km depths 365 velocity increase 293
hard porous 164 rock mechanics 5, 45, 75, 119, 227, 309, 424
heavily jointed 145 background, logic 191, 310, 390, 407, 418, 427, 434,
importance of rock type 3 442, 444, 463, 465, 502, 523
joints (ubiquitious term, see joint) 493 developments 197
mass (see rock mass) effect 192
massiv-, negligible jointing, QH2O – K estimation 558 engineers 189
matrix 30 experience 407
model, biaxial tests of fractures 422 modelling (see numerical) 133, 420, 427, 460, 461
non-brittle 304 rejection of elastic isotropic behaviour 382
non-dilatant 304 units (e.g. MPa/mm) 200
outcrops 11 wellbore (borehole) studies 302–311, 319
partially molten- 242 rock physics 45, 179, 192, 295, 350
quality 5, 88 at laboratory scale 323–367
qualities (A to F) and velocities (VP) 143 goals, exploration related (King) 323
quality Q (see Q-value) high pressure testing needs 323
skeleton 320 model for computing change of density, from VP and
salt, VP and density 326 VS 398
stimulant, unrealistic 417, 418 more ordered relationships of- 323
slope 30 reduced sensitivity for – at great stress 320
slope, modelled 517, 551 shallow perspective of- (see Ch. 2)
soft 93, 304 weathering and alteration in near-surface – (see Ch.2)
soft, massive 304 Rocky Mountains 231
soft plastic 304 roof fall 60
strength 3, 92, 150 room temperature 30
strength, reduced 309 Rose area, East Pacific rise 275
stresses 295–300 Ross ice shelf 253
Index 699

Ross Sea 253 bias, vertical wells, vertical structure 407


rotating S-wave transducer 356, 357 dense- near fault zones 394
rotation dense- near fracture zones 394
counter- of fluid lenses and rock-to-rock contact with shear sampling-induced microcracking 347
384, 527, 548–551 initial stress release when coring 347
roughness (see joint, fracture) San Andreas fault zone SAFZ 90, 91, 95, 221–224, 257,
profiles 508–510, 514, 545, 551 382, 429, 431, 432, 470
RQD rock quality designation 6, 7, 11, 12, 13, 16, 43, San Francisco Bay Area 257, 259
53, 69–72, 74–76, 78, 83, 88, 89, 92, 101, 105, San Joaquin Valley, southern California 260
106, 126, 144, 169, 175, 202, 225, 293, 447, sand 11, 12
448, 469 beach- 254, 350, 351
RQDo oriented 225, 411, 616 cemented-, model material 302, 303, 305
absence of- logging 208 distinguishing- and sandstone 351
RQD/Jn relative block size 469 fast- 15
scattering related 225 oil-, heavy 379
RQI reservoir quality index 333 sandstone, sandstones 12, 26, 31, 32, 94, 137,
link to FZI flow zone indicator 333, 335 325, 327
rule-of-thumb (i) sorting by name (see also Q and VP listings)
(Crampin) fracture density link to S-wave anisotropy 413, Bandera- VP static/VP dyn and Young’s modulus versus axial
415, 442 strain magnitude 340
particle size d95 x 4 ⬍ E aperture for groutability 487, Berea- 37, 58, 59
490 Berea- borehole studies 314, 315
Rulison Field 384 Berea- QP -confining and differential pressure, visible bed-
Russia, 88–90, 102, 117, 119, 170, 439 ding 191–193
RSR 76 Berea- QP -confining and differential pressure, frequency
187
S-wave (see shear wave) Berea- QP -differential pressure 352
anisotropy definition 438 Berea- QE -strain 188
lower propagation velocity 438 Berea- QE – % saturation, confinement 186, 187
particle motion 5, 438 Berea- QP and QS -differential pressure 342
phase velocities 480 Berea- VP static/VPdyn and Young’s modulus versus axial strain
pure S-wave, SS-wave 389, 438 magnitude 340
safety factor (F. of S.) 307 Berea- VP and QP -frequency -effective pressure, dry or brine
salt 12 saturated 344
bedded- 123 Boise- VP -VS -temperature 331
bedded, Ko 299 Boise- VP static/VPdyn and Young’s modulus versus axial strain
creep 321 magnitude 340
creep, dislocation climb 321 Boise- VP and QP -frequency -effective pressure, dry or brine
creep, undefined mechanism 321 saturated 344
-dome 380, 381 Castlegate, jointed 165, 167
Ko 299 CBTF well- VP – VS – density-permeability-birefringence
seal for hydrocarbons 295 445, 446
shear strength in situ 299 Crossland Hill- VP and VS versus 3D confining pressure,
-water (see brine) before/after fracturing, cubic specimens 362
weaker-, intolerance of stress difference 298 Crossland Hill- VS – permeability, before/after fracturing,
sample cubic specimens 364
boundary closure 193 Gypsy sands, tomography, QP and VP, also QP/VP
boundary opening 194 375–377
damage 516, 528–530 Kern River oil sand, VP-temperature, oil/gas % 330
disturbance, shale 354 Massilon- 1000/QE -velocity-strain magnitude 187
loaded to failure 195 Massilon- VP and QP -frequency -effective pressure, dry or
shortening, modulus 347 brine saturated 344
sampling (seismic) Michigan- QP – confining and differential pressure
bias, for rock physics testing 397 191–193
700 Index

sandstone, sandstones (contd ) brittle-ductile transition for- 312


North Sea- 3D qP-normalized permeability, stereograms -carbonate beds 387, 388
390 cataclastic flow of- 312
Pecos- VP static/VP dyn and Young’s modulus versus axial strain channel sand 375–377
magnitude 340 clay content in 24
Penrith- 1000/Q and Q versus 3D confining pressure, versus clay content %, VP –porosity 331–333
permeability, cubic specimens 363 clay-rich- 349
Ohio- VP static /VPdyn and Young’s modulus versus axial strain clay-rich- (reduced ␣-coefficient) 320
magnitude 340 clean, porosity-velocity, sand suspension 324
Troll sand 326 deeply buried 312
Venezuela oil sand, VP-temperature, oil/brine % 330 density and velocity data 326, 327
Whitchester- VP -pressure/depth, core and sonic log compari- epoxy-, model 470, 480, 481
son 348 fine clay layering 357–359
(ii) sorting by geological dating finely interlayered, distinguished from other rocks 379
Carboniferous 207–209, 346, 356 grain-crushing of coarse-grained- 311, 312
Cretaceous 231 hydrocarbon-bearing 370
Lower Jurassic 427 interbedded- 47, 145, 207–209, 212, 346–348, 450
Mesozoic- 153 joint, initial stiffness 494
Mississippian 343, 344 joint normal stiffness – stress 496
Permian- 72 joint normal stiffness, initial 495
Pliocene 343, 344 joint, stress-deformation 495
Triassic- 72, 427 Ko 299, 300
(iii) sorting by category lithified (sands) 374
porosity-velocity 24, 311 low porosity 31, 32
high porosity, velocity-porosity data 325 low matrix permeability 357–359
high porosity, velocity-density data 327 -marl sequence 46, 171, 172
high pressure-porosity tests 312 meta- 146
porosity-permeability 334, 343 mixed shale- units 369
porosity-permeability-FZI 335 -mudstone, interbedded 100
sandstone, sand, Tertiary data multiple clean/shaly sand composites, VP – saturation,
numerous engineering properties 19, 21 patchiness data 337, 338
porosity-permeability, core correlation 334 North Sea- 359, 360
well log, VP -core porosity-core permeability 334 Poissons ratio-pressure, dry, brine, crude oil 338, 339
sandstone, theoretical model porosity – VP (see listings following) 325
VP-pressure-brine/gas saturation 329 porosity-permeability (see listings following)
Poisson’s ratio-pressure-brine/gas saturation 329 QP – clay content and permeability 346
sandstone, tight gas QP -% compliant minerals, effective pressure 347
stress-strain data 339 QP and VP -effective pressure 347
3D, spherical, VP and permeability 357–359 Qseis (QP) magnitude only 181, 182, 186
lenticular, reservoir, VSP 471–475 QP/QS –(VP/VS)2 and pore pressure, to distinguish sand and
velocity-density data 327 sandstone 351
VP –VS data 328 1/Q – axial stress-strain, tuffaceous sandstone 195, 197
tight gas, 3D VP – confining pressure, spherical samples, 1000/QP and 1000/QE versus pressure, saturated 191, 192
layering/jointing, dry, saturated 358 1000/QE and 1000/QP confining pressure 191
tight gas, 3D permeability – confining pressure, spherical quartzitic 144, 146
samples, layering/jointing 359 reservoir 297
(iv) sandstone, diverse (alphabetic, see also Q and VP lists) 12, reservoir Q values 343
26, 31, 32, 94, 137, 325, 327 roughness profiles 509
artificial surfaces 80, 82 sand/sandstone 47, 324
aspect ratio influences for- 326 sand-shale proportions, each well 377
basic friction angles, dry, wet 511 saturated 12, 191, 192, 207–209
bedded and jointed- 207–209, 346–348 shear strength in situ 299
block of- 137, 545 ␴h min from minifrac testing 298
brittle- 306 stress-closure tests on joints 528
Index 701

stress-conducting aperture, CSTF 529, 530 attenuation, source of 306


stronger- tolerance of stress difference, shear stress losses (RQD/Jn related) 202, 225, 350
297–299 losses (Jr/Ja related) 350
strongly jointed 148 Rayleigh wave 196
suspension, sand 324 wave 191
tuffaceous Schlumberger 58, 206, 316
VP – % gas/brine 341 Schmidt hammer tests 26, 27
VP –pressure data, dry, saturated 32 in (TBM) tunnels 148, 149
VP –VS data 328, 329, 372 L-hammer 486, 509, 510
VP –VS data, varied effective stress 329 N-hammer 27, 28
VP/VS – differential pressure, also to extremely low rebound r and R 509, 510
pressure 351 schist 12, 32, 38, 100, 139–142
San Gabriel ranges 259 clay- 145
saprolite 4 (metamorphic) 212
Santa Barbara Channel, southern California 260 Pelona, S. California 261
saturated, saturation 8, 17 schistocity 3
complete 151 Schmidt net stereographic projection 358, 360
contra dry, joint stiffnesses 507–512 scour-holes, sediment-filled 56
degree of- (calculated) 126, 127 SCV, Stripa 133, 169
degree of- (measured) 21, 29, 135 sea 147
effect of- (gas/brine%, oil/gas%) 329, 330, 343, 345 floor interface 438
effect of- on VP 337, 338 -floor, rough 282
effect of- on VP, VP/VS and Qs/Qp cross-plot 189 water 15
effect of- on VP, QP for bedded-coal 201 seabed/sea bottom 56
effect of- on Kn and Kn dyn normal stiffnesses 202 cables (OBC) 391, 452
fully 15, 190 cable array, permanent 453
glycerol- (glass beads) 205 hydrophones 53, 56
heterogeneity with patchiness 336, 337 seismic 56
‘homogeneity’ with multiple units 336, 338 sealed fractures parallel H max are numerous 441
honey- Lucite plates 427 sealing
liquid-, high pore pressure 394 bulkheads 129–131
partially 182, 189 plastic layers 295, 297–299
partly- 5, 15, 17, 194 with hydrothermal minerals 270, 276
state of- of flat pores 190 with syn-kinematic, post-kinematic cements 441, 442
top of saturated zone 9 seasonal fluctuations
under-saturation 123 reservoir level 89
unsaturated 8 water level 89
vapour-, low pore pressure 394 sediment, sediments 3, 20, 205, 206, 221, 274
water- (ubiquitous term) 352, 353 Cretaceous 231
zone 9 deep accumulations of- 231, 232
saw-cut fractures, samples 70, 540 hard 249
S/C (ratio) subsidence/compaction Mesozoic 231
high- with discontinuum modelling 400, 454 newly deposited 296
low- with continuum modelling 400 post-rift (Atlantic margin) 255
scale soft 249
-effects (see joints) 507, 510, 522, 537, 538 thick-, effect on continental Lg coda 231
lack of-, assumed 400 unconsolidated- giving low Q 231
-length 373 unconsolidated- giving very low velocity 231
(scale dependent, see shear stiffness) sedimentary environment (ubiquitous term)
Scandinavian rocks 6, 13, 71 basin 179, 260
SCARABEE 110 diagenetic-based cycles 333
scattering (see also attenuation) 182, 195–197 ‘fining-up’ sorting technique 333–335
attenuation, calculated 197 layers, layering 14
attenuation, dual/triple poro-elastic models 461–481 stratigraphy matching 333
702 Index

sedimentary environment (ubiquitous term) (contd ) tomography (see tomography)


sedimentary rocks (ubiquitous term) 8, 25, 207, 263, transmission across joints 79, 80, 82
265, 299 wave blockage 231
Ko values 299 velocity (ubiquitous term, see VP, VS and data lists)
Q and Q⫺1 values to 2.4 km depth 232 velocity changes due to tunnelling 615
Q versus interval stacking velocity 234, 235 vessel 394, 396
sedimentation 300 wave scattering (see scattering) 231
-erosion, effects on E-modulus hypothesis velocity gradients (see VP-depth) 107, 246
saturated –rocks 207 seismically visible 3, 296
Segunda Angostura dam site 163 seismicity (ubiquitous term)
seismic base of- thermally controlled 257
anisotropy (see anisotropy) 40, 41, 42, 394 seismogram, seismograms
anomalies 4 earthquake- 213
attenuation (see attenuation) plane-layer- models 231
attenuation as sensitive indicator 386 subtraction of-, horizontal well sections 382
attenuation tomography 281 synthetic 231
coda waves (from earthquakes) 209–219 synthetic- not correlating with measured 317
-data, variation with offset and azimuth 382–396, synthetic- modelling 266
401–406, 407–482 (Ch 15) three-component 437
detection of subsidence 400 three-component, before/after rotation, fast/slow 437
disappearance 24 seismology
fissurization index K 70 broad-band 207–209
global measurement locations 246, 251 seismometers
high resolution- reflection 254 borehole 221–224
impedence (see impedence) in-well, in-borehole 209, 221–224
method, shortcomings of 14 wide band-width- 221–224
modelling (numerical) 204 selective firing
profile 7, 9, 11, 12 Sellafield site, N.W. England, UK Nirex Ltd. 57, 94, 224,
profile, continuous 283 309
profile, high-resolution 282 Rock Characterization Facility (planned) 528–530
profiling (horizontal HSP) 153–155 sequence shear zones 3
profiling (tunnel TSP) 153–155 sequential firing in perpendicular directions 313
processing (ubiquitous term) serpentinite 23
Q (ubiquitous term) Severn Estuary, second crossing 11
Q and similarity to Qrock when deep, intact 365 SH-wave 354, 357
Q , low values with fracturing 365–367 shaft 87, 121, 123
Q , low values with jointing 222, 223 deep 93
quality (inverse of attenuation) 17, 65, 127, 181–239 erosion 156
quality factor Q and components (see Q , Ch 10, and Ch 13) shale, shales 12, 27, 32, 94, 146
reflection (see reflection) 254 Antelope-, fractured 236, 237, 387, 388
reflection tomography (see tomography) attenuation in- 354–356
refraction method, survey, profiles (ubiquitous term) 4, 10, Brown-, sigmoidal fractured 236, 237, 387, 388
13, 14, 70, 76, 115, 155, 166, 173, 245–290 bulk-, not matching Qseis Emass model 387, 388
refraction, deep (0–35 km) 229 CBTF well- VP – VS – density-permeability-birefringence
response, temporal 307 445, 446
risk mitigation 189 clay-particles in- 418
shallow refraction- 3, 4, 9, 76, 115, 176 compacting 301
shallow refraction, in tunnel 122, 153 Cretaceous 231
shallow refraction, beach sand 253 dense 38, 39
sonde 151 finely interlayered, distinguished from other rocks 379
sources 189 fine layering in-, TIV symmetry 374
spectroscopy 118 ‘fining-up’ sorting 333
spread 140 Fiqua, Oman 456, 457
survey (ubiquitous term) interbedded 47, 234, 298, 300
Index 703

kerogen rich-, state of maturation 379 fracture microseismicity, Ekofisk 443


Ko 299, 300 fracturing, mini-EDZ, fluctuating sonic velocities 379
Kimmeridge Bay, Dorset 488 log-spiral- 302–307
Kimmeridge-, North Sea 351–353 micro shearing 282
-limestone interbeds 148, 234 modulus and fluid type 420
low permeability- 268 on bedding planes 304
Mancos 167 permeability, effect on 517–536
mixed-sandstone units 369 relaxation 187
mud-filtrate example 317 slowness 316
overlying- 393 strain 189
over-pressured, lab test 354–356 tests, laboratory
Palaezoic 430 tests, in situ/field 20, 101
Qseis 181, 182 zones 9
Reskageage, Iceland 504 shear modulus (␮) (see also deformation) 5, 13, 71, 104,
-rich layers 376 109, 111, 161
-sandstone units 370–372 pressure-sensitive-, using excess compliance 397
saturated 12 shear stiffness KS (joints, fractures) 45, 346
sealing, caprock 295, 298–300 clay-filled discontinuities, normal stress 505
shear strength in situ 299 equations 511
␴h min from minifrac testing 298 dynamic- (see dynamic) 282, 423, 424
-smear sealing 372 pseudo-static 282, 422, 492–499
stress-conducting aperture, CSTF 529, 530 reductions with block size 483, 484, 523, 524
thin-bedded 300 reductions with stick-slip, assumed 435
Tournemire tunnel EDZ in- 121 scale (block-size) dependent- 6, 492, 493, 501
VP and density 326 scale effects discussion 421, 422
VP –VS data 372 shear strength
weaker-, intolerance of stress difference 298 displacement dilation curves 507, 510, 526
shale alteration 302, 312, 313, 317 displacement dilation modelling, Barton-Bandis
shallow 506, 507
depth 147 equations 511
gas (see gas) loss of- 241
layer velocities 140, 156 low- of suspension 350
sites 7, 74 peak- of joints 491, 507, 510–512
water flows SWF 350 peak- to match mobilized friction, conductive fractures
shear, shearing 544, 545
and dilation 390 pre-peak or post-peak friction mobilization 545–548
body-waves 5 residual- of joints 491, 507, 510–512
box (DST) 510 shear strength envelope for
box samples 524 filled (clay-) discontinuities 538
causing dilation 249, 258, 259 intact rock, high stress 538, 540
causing maintenance of permeability 258, 259 intact rock, numerous data, high stress 539
compliance 346 induced fractures, extreme (tectonophysics) stress 537, 538
deformation zones 308 induced fractures, high stress 538
direct- tests (DST) 401 rock joints, lab and in situ, engineering stress 537, 538
direction 39 rockfill/crushed rock 538
displacement 9, 84 sandstones, high stress 312
displacements to peak, joints, discontinuities, many shear stress 5, 52, 165
scales 546 change of 217
displacements, and stick-slip 435 displacement 507, 510
down-dip-, compaction mechanism 392 dissipation, through log-spiral failure 306
failure 9, 130 limiting value of 227
failure of matrix during water-flooding 403 maximized 308
failure of model rock masses 105 -normal stress envelopes, induced fractures, extreme stress
failure surfaces observed 305 537, 538
704 Index

shear stress (contd) velocity (VS) examples (laboratory measurements/rock


-normal stress envelopes, joints: lab and field, engineering physics) 6, 7, 8, 29, 32, 36, 37, 81, 87, 109, 131,
stress levels 537, 538 136, 189, 265, 298, 328, 329, 331, 351, 353, 354,
resistance to 227 364, 372
effects on velocity 40 velocity (VS) examples (field surveys, all depths) 6, 7, 8, 9,
shear wave, waves 7, 32, 36, 37, 81 12, 32, 64, 70, 71, 75, 92, 92, 90, 104, 105, 111,
ability to penetrate gas cloud 391 161, 229, 241, 249, 250, 251–253, 264, 266, 267,
amplitude 110 270, 275, 277, 280, 281, 286, 379, 395
amplitude decrease 350 shear wave splitting (and polarization) 82, 179, 313,
amplitude, effect of stick-slip, sliding 435 314, 354, 355, 357, 369, 372, 386, 388, 389,
anisotropy 9 407–481
anisotropy and fluid type, gas, brine 456, 457 above sedimentary hydrocarbon basins 408
anisotropy and fracture dip 445 above earthquakes 428–438
anisotropy % linked to crack density 413 above small earthquakes 408
anisotropy %, fracture porosity, gas or brine 457 anisotropy parameter 451
anisotropy % linked to pore pressure, 90°-flip model 416 anisotropy parameter range 451
anisotropy % linked to frequency and fracture sizes 467, argillaceous rocks 450–452
470, 472 as function of (joint, fracture) shearing 518–520, 522
anisotropy linked to permeability 445–446 correlated with subsidence bowl 412
anisotropy logging 302, 312–316 due to component sets 386
anisotropy sources (see anisotropy) 409 due to conjugate sets 483
converted from P-S waves 391 due to high ZT, low KS supposition 512
detection of oil or gas, compressibility re ZN 407 due to intra-bed joint stretch (supposition) 453
energy 130 due to microcracks or joints, discussion 432
fast- with longer wave length 314 due to ‘stress or strain’ 453
fast- parallel to formation fast axis 313, 314, 417 due to subsidence, bed stretch, intra-bed jointing (supposi-
fast- parallel to structure 313, 314 tion) 451–453
flexural- 312–316 due to sub-vertical fractures 450
frequency 81, 110, 112 focussed on structural domain with VSP 410
frequency anisotropy 459 influenced by sub-recording station structure with earth-
leading split- stable, fast direction 411 quakes 410
monitoring of fracture closure 364 matrix with relevant compliances 417
monitoring of fracture closure cycles 364 mechanism explained 439
polarization 82, 313, 357, 359, 361, 445, 446, 448–477 Mid-Atlantic Ridge 437
polarization affected by shear stress, shearing, speculation multiple- 410
517–521, 525–527 Natih field, Oman 455–458
polarization in principal stress direction 359, 361 New Madrid seismic zone 428, 429, 430
polarized- 393 non-parallel ␴H max direction 384, 385, 389, 390, 403,
processing, demands of 438 404, 406, 414, 429, 431, 432
slow- perpendicular formation fast axis 314, 417 parallel ␴H max direction 429, 436, 437, 439, 446, 447,
slow- (perpendicular to structure) 314, 408 448, 450
slow-, with gas in fractures 456 Parkfield seismic monitoring array 429, 431, 432
source 439 petroleum reservoirs 438–440, 442–460
splitting (see shear-wave splitting) polarization examples 430, 431–433, 437
surveys 12 polarization in anisotropic zone 439
technology, belated application, mature reservoir 452 shallow, 15–30 m 447, 448
teleseismic- 410 shallow, subsided overburden, Valhall 453, 454
three-component sensors, 1  v, 2  h 438, 439 sources of- 410
(with hydrophone 4C) 438, 439 temporal changes 410
time-delay, temporal 433 using near-offset VSP 445
travel-time, temporal 433 Valhall overburden 453, 454
velocity anisotropy 408 shear wave window
velocity transition 231 above earthquakes 409, 428, 436, 437
velocity (less than fluid velocity) 206 epicentral distance less than focal depth 409
Index 705

Shell 456–458 slickensides


Shetland Islands 404 on conjugate joints under compaction, Ekofisk 443, 452
shields (geologic structures) 246 sliding
Canadian 217 on crack faces 182, 190
Indian 219, 230 stable- on fracture surface 540
shield area 217 slip
Shinkansen high speed railway, Japan 47 coupling of normal and tangential- 420
shooting vessel/boat 396 on conjugate joint sets 308, 420
shortest path on pre-existing faults 130
through best rock 14 slope
shotcrete 80 reinforcement 117
steel-fibre reinforced 621 slow
Siberian Shield 230 compressional wave 209
silicate host rocks 23 direction 128
silica shear wave depends on fluid in fractures 457, 458
fume (see micro-silica) smectite 352
Qseis 181 SMS stress monitoring site, Iceland 432, 433
Siljan Ring, Sweden, borehole data 543 Snow 3D network model 172, 173, 487, 490
silty flood-plane 375 Snell’s law 438
siltstones 26, 146 soft (porous) rocks 48, 79, 101, 103, 113
above shale, caprock 454, 455 softening behaviour 303
basic friction angles, dry, wet 511 soil 9, 11
Carboniferous 207–209, 234, 346–348, 356, 357 clayey 12
inter-bedded 207–209, 234, 346–348 engineering 106
Ko 299, 300 silty 12
joint, initial stiffness 494 to clay 24
joint, stress-deformation 495 vegetation 24
joint normal stiffness – stress 496 solution channels 378
joint normal stiffness, initial 495 sonar buoys 11
layered sequence 234 sonobuoys 261, 275
stress-closure tests on joints 528 sonic
Tertiary 304 and ultrasonic tests 191
Triassic, North Sea 427 log 45–47, 52, 69, 234, 271, 301
Lower Jurassic, North Sea 427 log, core correlation 277
single log fluctuations 301, 310
-bit run 312, 313 logging 380, 386, 387
singularity logging of boreholes 97, 119, 129, 207–209, 261
P-wave cancellation 466 logging tool 313
zero attenuation 466 probe 129
sinusoidal shear and compressional- log 301
shear strain 5 Soultz deep borehole data 555
Site 977, ODP Leg 161 (Shipboard Scientific Party) 334 soundings 9
SKB 131–136, 203 source, sources
Skien river, Norway 161, 162 -and receiver (pairs) 204, 205
skarn ore 65 at the tunnel face 154
slate, slates 12, 38 calibration shots 394
cleavage fractures 362 lines (3D-4C) 391
cleavage, stress-closure 494 micro-earthquakes 209–232, 394
joint normal stiffness, initial 495 multiple, fixed offset 440
joint, stress-deformation 495 of error 169
Mesozoic 153 on the surface 154
high-grade 248 P- and two orthogonal S-, nine component 471
stress-closure tests on joints 528 quarry blasts as- 394
strongly jointed 148 receiver lines, three for strike detection 446
706 Index

source, sources (contd ) steam-flooding 329


S (SS)-wave source, sea-bed suction anchor 438 causing local heating 329
synchronise- 153 steel
separation of- 9 indentor 59
vibrator, 1000 positions/day 455 Qseis 181
source rocks 372 steeply-dipping
South Africa 110, 112, 131 faults 257
South America 230, 231 joints (see vertical)
South Carolina 90, 92 planes 257
South China Sea, sonic log of reservoir 301, 313 stick-slip 110
South Dakota 60 on saw-cut 540
South Korea 109, 111, 169 onset of- 540
Southern California 228, 258 stiffness (see joint, fracture, normal, shear)
spacing (see joints, fractures) stiffness (of joint or fracture)
span (of tunnel or cavern) 151 anisotropy Kn dyn  Ks dyn 518, 519
sparker 67 -compliance comparisons 418–425, 427, 428
spatial resolution 382 data gaps 424, 425
of variable structure 382 dynamic and static (Kn) comparison 422–424
of temporal changes 382 -fracture flow relation 425
specific matrices 374, 477
stiffness, pseudo-static (see also normal stiffness) 198 mechanical 169
spectral amplitudes/magnitudes 198, 199 normal Kn 6, 17, 198, 199, 202, 282, 418,
ratios 199, 212 421–424
ratio method (see Q analysis) 353, 362 normal Kn greater with stiffer fluid 420
split shear wave (see shear wave splitting) normal Kn much greater than Ks 550
spreading mid-ocean ridge (see mid-) inequality of Kn and Ks 421–424
spherical real and imaginary parts 475
hemi- projection 358, 360, 445 ratios of Kn/Ks 418, 421
samples 357–360 ratios of Kn/Ks, model fractures, prototype stress 552
squeezing (see tunnels, boreholes) 297, 299 ratios of Kn/Ks interpreted for fault zone 427
squirt ratios of Kn/Ks, in situ, saturated, weak rock 512
clay- flow 346 relations to ZN and ZT compliances 418
flow 17, 183, 187, 206, 346, 349, 457, 461, shear- Ks 6, 418, 428
465–476 shear- Ks for large scale features, faults 427, 428
flow in poro-elastic models 461–481 specific- of joints 422
flow absence when dry 341 welded asperity Lucite-laminate assumption 418
flow absence when saturated 341 Stirling Castle 118
flow mechanism 384 stochastic simulation
flow reduced by pressure, closure 346 of oil saturation changes 398
flow related to loss in bulk modulus 356 Stone Canyon earthquakes, California 210, 211
flow related to small loss in shear modulus 356 Stoneley wave 316
losses (Jr, Ja related) 202 Straight Creek Tunnel, Colorado, USA
phenomena 182, 190 pilot bore 139–142
SRF stress reduction factor (Q-parameter) 92, 175, 225, strain 4
293, 350, 447, 448, App. A, 615–625 amplitude 183
stability amplitude, axial 340
poor 144, 145, 149 amplitude, importance of 183, 187–189, 193, 194
steam amplitude-frequency plot 183
driven TBM 319 axial- (see also axial) 339
-flood larger- territory 183
injection 66 microstrain (see microstrain)
injection cycle 379 radial 339
infection-front imaging 378, 379 tangential 304, 305, 307
Index 707

uniaxial 5 -induced failure, tunnels 304


volumetric 5 -induced fracturing 127
strength (see also rock, joint, fracture, compressive, -induced joint closure 13
uniaxial) 27 level 3, 16, 30
corrosion 304 low horizontal- 367
deformation components 520, 522, 523 locked-in- (grouting) 172
high 22 maximum horizontal- ␴H max 88, 382, 429, 472
post-peak- loss 304 maximum horizontal- ␴H max direction re water or
resisting critical crack density worries 413 not 384
strengthening-by-confinement 537–539 minimum 296–300
stress (pressure in some contexts) minimum horizontal- ␴h min 88, 298–300, 406
aligned, fluid-filled cracks 409 mis-alignment with major- 506
anisotropy 9, 115, 297–300 monitoring site, Iceland 432–434
anisotropy, rotated 160 negative effective- 278
axial 339 normal (joint or fracture) 79, 198, 199
azimuth, as fracture azimuth? 382–386, 388–396, principal- 115, 128, 132, 136, 137
401–406, 407–482 principal- directions 137, 313, 315, 409
changes due to tunnelling 615 principal- directions, above/below shearing joint
closure tests on joints 198, 485, 487, 494, 495 519–521
confining-, excessively high 265 principal- directions, assuming horiz./vert. 409.
confining-, pressure 248, 250, 257, 263, 265 principal-, modelled 309, 310
deformation loading of joints 422–424 radial 30, 120, 615
deformation loading of jointed rock 484–486 radial stress release 123
deformation loops 97, 98 redistribution (ubiquitous) 30, 125
deformation gradients 184 sensitivity assumption, collective 397
dependent velocity 35 shear- (see shear stress)
deviatoric- contours 131 -slabbing, thin-walled 304
deformation flow monitoring, URL fracture zone 505, -strain curves 37, 195, 300
506 -strain loops, cusped to elliptical 188
difference 297–299 strength ratios (␴n /JCS), CSFT tests 528–530
difference intolerance by weak rock 297–299, 302–304 tangential stress 30, 93, 115, 120, 122, 123, 127, 129,
differential- 37, 268 132, 136–138
differential- contra effective- 356 tangential stress concentration 301, 379
discontinuity, re splitting 518–520 tangential stress close to wall 310, 615
-displacement behaviour (joints, fractures) 87 tangential stress compaction effects 156, 301, 379
dissipation into rock mass 305 tangential stress enhanced properties 316
distribution 129 tangential stress maxima and minima 306
effective- 88–90, 263, 266, 267, 269, 295–297 tangential stress components of similar magnitude 311
effective- coefficient (Biot) 191 thermally-induced- 135
effective-, extremely low 350, 351 3D stress state (ubiquitous state) 302
effective- gradients 267 total 298
effective- increase prior to water flood 401 transfer (across joints) 172
effective- reduction as result of cold water injection 414 vertical
-gradients 60, 61 ␴H max direction
-gradients, radial 314, 315 bisected/intersected by 390
high- 22 bisecting water-flood directionality 403
high- gradients 61 comparison with fractures 404
high- monitoring 60 non-alignment with- 527
high- region 59 stressing
horizontal- 128, 135 post- (effect of grouting) 170
horizontal, below quarry floor 500 stress transformation equations 552
horizontal- enhancement, over magma-chamber 282 error for all dilating geotechnical materials 554
indentor- 59 error for dense rockfill, sand, OC-clay 554
708 Index

stress transformation equations (contd ) survey ships 243


error for non-planar rock joints, fractures 554 suspension
stress transformation equations, modified 553 sand in state of- 324, 350, 351
Stripa Mine, Sweden 119, 133–136, 197–200, 422–424 Sweden 77, 119, 197, 203, 543, 555
structural, structurally Switzerland 38, 76, 124, 161, 169
controlled fall-out 319, 320 symmetry
geology (see geology) deviation from higher- 360
orientation data integrated 384, 385 directions defined, MON, ORT, TI 360
St.Venant principle, violation monoclinic MON 357, 358, 360
non-coaxial stress and strain due to dilation 552 orthorhombic ORT 360
subduction zone, oceanic 179, 226–228 plane 440
sub plane of circular- 355, 356
-fjord 56 transversely isotropic TI 360
-sea link 320 vertical- axis 45, 372
-sea sediments 9, 56 synthetic sandstone crack model 480, 481
-surface 3, 4, 7
-surface interfaces 438, 439 Taiwan 205
-surface resolution, improvement 378 tangent and secant slopes/gradients 98, 101, 198
-surface topography 4 tangential stress, see stress
sub-ocean, sub-oceanic Tangsham earthquake, China 110
floor attenuation 268, 269, 281, 282, 286 Tanzania 72
Layer 1, 2, 3 sub-ocean divisions 243, 244 target
Layer 2A, 2B, 2C 277, 283–285 horizon 15
Layer 4, 5, 6, sub-ocean divisions 244 fracture zone 389, 391
spreading ridge velocity modelling 261–294, 317 tar sands
velocity structures 262–290 steam flood in- 331
subsidence bowl 453 reduced viscosity close to wells 331
match to continuum modelling 400 VP reduction with heating 331
match to discontinuum modelling 400 TBM (see tunnel boring machine)
match to shear wave splitting 453 tunnel EDZ, modelled 309, 310
subsidence/compaction S/C (ratio) JRC logging in – tunnel 543
high- with discontinuum modelling 400, 454 Technical University of Trondheim 457
low- with continuum modelling 400 tectonic stresses (ubiquitous term)
submersible 266, 273 reginal- at San Andreas fault zone 432
subway station 4 tectonophysicists 210, 246
Sudbury, Ontario 61 televiewer (see borehole)
sugar-cube model representation 479, 480 temperature
Sugar-Loaf, Rio de Janeiro, estimated Q, QP, K 558 anomalies, conductive fractures 541, 542
sulphide orebody 61, 62 applied to HTM block test 514–517
superficial deposits 12 corrected 256
superimposed multiple ellipses effect on resistivity 297
at faults, re elliptical Vfast and Vslow 394 effects on joint apertures 198
super-k poroelastic model 377 effects on VP of oil sands 330
support pressure (tunnels, caverns) 92, 118 effects on VP and Poisson’s ratio of oil sand in situ 379
supposition elevated- testing needs 397
influence of ‘average rock’ (earthquake source) 442 gradient in crust 242
influence of ‘biased sample’ (fractured reservoir) 442 low- 33
O/R contrary reflection/polarization rotations 414, 446, -VP of oil sands in presence of oil/gas/brine 330
450, 453, 474 -VP and VS , Boise sandstone 331
surface temporal variation of
ejection of water and sand 219 anisotropy 454, 455
liquefaction 219 attenuation 413, 414
-magnified ground deformation 219 attenuation, differential 454, 455
outcrop joint orientations 384–386 GPS displacements 413
Index 709

polarization directions 413, 414, 452, 453 Barton,Colleen, Zoback, Moos,Townend, conducting joints
pressure in well 433 under shear stress 541–543
qS1 reversal with qS2 454, 455 Barton, e and E apertures from JRC 364, 489
seismic events (before/after earthquakes) 215, 216, 218 Barton, JRC-mobilized concept 435, 491, 548
S-wave time-delays 429, 433 Barton, M and Qseis similarity for jointed rock masses
S-wave travel-times 433 202, 203, 210, 220, 221, 224–226, 269, 348, 350,
tensile 352, 365–367, 387, 405, 410, 411, 424, 436, 476,
fracture traces 385 499, 500
strength (see rock) Barton, natural block-size scaling of Ks 422
tension, tensile Barton, non-linear strength envelopes, JRC, R, log- formula-
fractures 81, 87, 299 tions 538
fractures, models 524, 527, 548, 551, 552 Barton 90°-flip ␴h min to ␴Hmax reversal, alternative
hydraulic- 299 theory 416
tensor Barton, Qrock to Lugeon inversion 175, 176, 366, 555,
elements 374 556
fourth and second rank- 417–419 Barton, Q-system of rock mass classification 615–624
geometric 174, 292 Barton, QH2O permeability estimation, depth dependent
hydraulic 174, 292 556–558
Terlingham Tunnel 94. Barton, roughness profile, tilt test for JRC 508–510, 514
terrain Barton, shear stiffness scale effect 493
mountainous- 33, 56 Barton, stress transformation equations with dilation 552,
steep 56 553
TerraTek, Salt Lake City (now Schlumberger) 509, Barton, thermal over-closure of joints 513–517
513–515, 518, 552, 553 Barton, VP Qc M linkage 146
Tertiary rocks 142 BISQ Biot and squirt-flow, Dvorkin 378
granites 19 BOSK saturated, fractured medium model 457
mudstones 19, 27, 304 Crampin EDA shear-wave anisotropy, basic theory 429
oil shale Ko 299 Crampin 90°-flip excess pore-pressure theory 416
rocks 142 extended-Kuster-Toksöz, velocity-crack aspect ratio model-
sandstones 27, 144 ling 276
sandstones Ko 299 extended-Walsh, velocity-crack aspect ratio modelling 276
sandstone-shale units 370–372 Gassmann (Biot-) fluid substitution for porous media 317,
sedimentary rocks 25 420
siltstones 304 (error for fractured porous rock) 420, 457, 458
shale Ko 299 Johnston, spectral amplitudes, dispersive waves 198
tuff Ko 299 Jones, dispersive squirt flow model 208
Terzaghi, theory of effective stress 268, 320 Hudson (seismic) effective cracked-media model 196, 461
test apparatus, equipment (diagrams) Hudson (seismic) normal and shear compliance model 427,
rock physics 59, 192, 197, 361 503
Texaco 374 Hudson and Crampin P-wave anisotropy, fracture density
Texas 203, 386 503, 504
Texas, fractured carbonate gas field 536 Hudson and Schoenberg compliance formulation for faults
Theories reviewed (brief, biased selection, see references) 504
APE, Zatzepin, Crampin 416 Liu, fluid-type, fracture compliance, aspect ratio
Aki conjecture 228 theory 426
Aki and Richards crack relaxation mechanism 387 Mohr-Coulomb shear strength criterion 461
Bandis hyperbolic joint closure model 428, 485, 486, 494, Nur, velocity anisotropy due to loaded micro-cracks 36
528 Nur, critical porosity 324
Bandis shear strength scale effects 522 Rüger reflection coefficients with TIH 390
Barton-Bandis joint constitutive laws 435, 511 Schoenberg, slip-interface, displacement discontinuity 200,
Barton-Bandis strength-displacement-dilation, scaled 487, 461, 501
493, 507, 509, 511, 526, 531–535 Schoenberg-Sayers excess compliance theory 416, 417, 503,
Barton-Choubey peak shear strength, JCS, JRC, stiffness 504
485, 507–512, 514 Stratification-percolation model for apertures 498
710 Index

Theories reviewed (brief, biased selection, see delay increase in reservoir formation 471, 473
references) (contd ) delay-depth intervals: anisotropic, isotropic, anisotropic
Tod crack density decay model 461 473
Walsh elliptical crack closure model 364, 529 delay variations 445
Walsh friction dissipation in micro-cracks 182 distance inversion method 275
Zoback hydrostatic not lithostatic pore pressure in crust distortion 14
(references) average equation 30, 119, 121, 323, 324
thermal lapse comparison 399
gradient 242 lapse cross-well data 379
over-closure (see over-closure of joints) lapse survey 66, 451
relaxation 183 lapse, time shifts, hysteresis, in-outside reservoir 484
thermally-induced stress 135 separation, increase with path length 439
thermal expansion sag, gas or over-pressure 381
cofficient of contained water 321 shift in compacting reservoir 398, 399, 402
coefficient of contained fluids 351 TIV 374, 390, 394, 451
coefficient, importance of included joint 516 tomograms
effect on crustal velocities 248–250 acoustic 59
thin bed alternative 64, 65
in relation to wave length 372 amplitude attenuation- 65
Thomsen, weakly transversely isotropic 356 difference- 63
1000/QP and Q scale comparison 380 error- 63
3D fence- 64
finite difference modelling 196 for blast monitoring 61, 63, 64
measurements (axis-by-axis) 354–357, 359–364 non-co-planar- 64
multi-azimuth walk-away 369 Poisson’s ratio- 64
full-azimuth, full-offset P-wave survey 526 Pulse broadening- 65
multi-component, multi-mode, multi-azimuth S-wave- 64
acquisition 382 2D and 3D comparison 259
P-wave velocity tomograms 375 time-lapse S-wave-, stable through fluid change 379
repeated- multi-component survey 453 velocity difference- 61, 63
seismic survey 399 velocity difference- at fault zone 394
spherical sample measurements 358–360 VP-, and Poisson’s ratio- 379
3DEC 133, 365, 454 VP and QP-, depth slices 375
3D-4C (OBC data acquisition) 391, 396 VP/VS- 64
4D (see four) tomographic, tomography (cross-hole seismic) 40, 49, 52–65,
three 92, 94, 131, 132, 139, 154, 156, 169, 257,
-component sources 7 374–378
-dimensional failure surfaces 306 amplitude- SAT 83
-dimensional P-wave measurements 357–360 back-projection- 231
-dimensional seismic (ubiquitous term) 281, 282 blasting effects 63–64
thrust fault 244 attenuation- 18
Tibetan Plateau, China (thickest crust) 245 cross-continental 230, 231
Tier I, II, III classification (see HPHT) 320, 321 cross-hole- 52–58, 61, 74, 88
tight (fractured) gas cross-hole- (deep boreholes) 57, 94, 115
reservoir 384, 471–475 cross-pillar- 60, 61
sandstones 357 cross-well- 233
sandstones, lenticular 471–475 cross-well-, with permeability 377, 378
TIH (also HTI) 374, 390, 394, 417, 477 -inversion 56
tilt test for JRC estimation (see joint index tests) 514 post-blast 63, 64
time pre-blast 63, 64
delay (between split shear waves) 313, 314, 409, 412, 398, radar attenuation difference- 169
399, 429, 432–434, 436, 437 radar slowness- 169
delay from micro-earthquakes (Iceland, Hawaii, San shallow 3D- 374, 375
Andreas Fault) 436 3D velocity- 259
Index 711

3D attenuation-, sub-ocean floor basalts, magma chamber tropical terrains 4


281, 282 Tsukuba Oishiyama earthquakes, Japan 210
time-lapse imaging 60 tsunami waves 404
velocity- SVT 83, 169 tube waves 44
Tonga Ridge 179, 227 tuff 26
Tonga Trench 179, 227 breccia 26
tool x- and y-axis 313 jointed, numerical model 309
topography roughness profiles 509, 545
exact 15 stress-conducting aperture, CSTF 529, 530
Tor Formation 391 Tertiary 211
Tournemire experimental tunnel, Aveyron, France 39 welded 94, 510
Tower Colliery, Wollongong, Australia 200 tunnel, tunnels 3, 25, 30, 52–56, 80, 93, 94, 224, 339,
Tracer injection 359 413, 615–621
transducer ahead of- 151–156
pressure 174 arch 4, 141
temperature 174 boring machine TBM 4, 17, 24, 131–133, 140,
transient crater 404 148–157, 321
transition boring machine TBM back-up 312
dry-saturated- 14 by-pass- 56
transmitter-receiver 62 collapse 4
distance 120 data (VP – UCS, weak rock) 25
sonde 125 deep 24
Trans Manche Link 94 deformation (see also displacement) 173
transmissivity 161, 492 deformation back-analysis 192
shear displacement effect on- 492 difficulties, major 156
transversely isotopic (see TIH, TIV) engineering 301
travel experimental- 117, 118, 124–133
distance 30 exploratory- 107, 119
inversion 15 exploratory boreholes 46, 109, 111
path length 198 face 139
travel time 15 face, ahead of the- 153
difference, fast and slow S-waves 408 face collapse 153
-distance-depth plots 14 face-log 94
increase perpendicular to failure zone 540 failure modes, multiple 304
inversion, orientation 404 hydropower 120–122, 148–150
post-stack changes of- 398 in-tunnel reflection (see seismic profiling, HSP, TSP)
pre-stack changes of- 398 logging 615–626
two-way- 447 measurements 71, 80, 120–133, 139–158
Trawsfynydd nuclear power station 9 models 306, 307–311
Triassic 447 mountain 56, 111, 153–157
crystalline basement 220 pilot 53
mudstone 11 pipeline 146
sandstone 11, 72 pressure 300, 536
siltstone 11 rail 47, 48, 141–143, 408
triaxial (see stress) reinforcement and support 93, 141–143, 145–149, 152,
cell, conventional 362 156
confinement 192 road (see also sub-sea) 53, 54, 56, 146
compression 343 sealing experiment 129–131
ship-board 263, 269 stability despite ‘critical’ crack density 413
stress 267 sub-sea road- 56, 139, 144–147
tests, high pressure 312 support 140–142, 143, 145–149, 151, 152
triple-porosity, matrix, micro-, meso-fractures 465–474 wall 30
Troll sand, North Sea 326 water supply 72–73, 146
Troodos ophiolite, Cyprus 278 wedge release 301
712 Index

tunnelling 4, 48, 52–56, 109, 161, 444 under-pressured 351


in squeezing rock 306 - and over-pressured 351
rock quality variations in- 394 underground research laboratory URL, Manitoba, Canada
turbidite 9 57, 87, 117, 127–131, 505
2D two-dimensional models 307–310, 369 fracture zone study 505, 506
Tydeman North Atlantic fracture zone profile 273 undisturbed medium 5
Type A, B, C rock mass load-deformation 485 undrained
Type I, II, III, IV reservoirs 373 shear strength 303
type curves 164 undrained behaviour assumed
Tytherington quarry, Bristol, SW England 500, 503 high Q data-anomalies for metasediments 234
low Q⫺1 data-anomalies for sub-ocean basalts 269
UCS (uniaxial compressive strength, ␴c ) 70, 92, 93, 144, unexpected events (TBM tunnelling) 158
150, 164, 225, 263, 305–307, 320, 365, 366, 448, uniaxial (see also one-dimensional)
449, 486, 507, 510, 556 compressive strength (see UCS)
effect of moisture 511 compression testing (see also VP -) 23, 25, 26–28, 195
UDEC 365 cycling 339
attenuation modelling , fractured 478, 479 loading of in situ blocks 165–167, 513, 514
joint interactions in- 393 loading of borehole test blocks 137, 315
-MC modelling of tunnels, boreholes 308 strain testing, of core plugs 400
-MC modelling of subsidence 400, 401, 454 University of Berkeley 197–200, 464
-MC 306, 308 University of Cergy-Pontoise 464
UDEC-BB modelling University of Edinburgh 450–452, 464
of compaction 400, 401, 443, 452 units
of compaction, displacing secondary set 550 compliance-stiffness discussion 423, 424, 487
of jointed rock masses 484–486 friendly-, MPa/micron 500
of HM(T) block test 518, 519 recognisable- of stiffness 189, 202, 421
of tunnels, boreholes 133, 306, 309, 310, 467, 469 resemblance to- of GPa (see QP similarity) 191,
in rock mechanics 484, 508 192, 200–202, 203, 210, 220, 221,
UK 16, 57, 404 224–226, 232, 269, 348, 350, 352,
-France 17 365–367, 377, 379, 387, 405, 410, 411,
Nirex Ltd. 57, 94, 224, 309, 310, 387, 528 499, 500
ultimate strength 37 unloaded 30
ultrafine cement 170 unloading 30, 97, 106
ultramafic rocks 12 unmated fractures/joints 86, 87
ultrasonic up-hole shooting 14
frequencies (ubiquitous term in Ch 13) 269, Upper Cretaceous chalk 447
339, 340 Upper Permian 447
frequency, relevance to sonic, seismic exploration 349 upper crust, New Madrid seismic zone 430
logging 120 upper mantle 228, 241
P- and S-wave components 355, 361, 362 anisotropy 276
measurements 44 anisotropy due to aligned olivine 276
pulse echo technique 356, 357 crustal velocities (see sub-ocean) 283
pulse transmission 38 high- temperatures 231
pulse transmission equipment 192 Ural Mountains 230
reflection technique 355, 356 URL, Canada (see underground)
stress sensitivity with-, less than with seismic 397 USA, United States of America 42, 60, 90–92, 102, 123,
testing 372 165–167, 203, 210–216, 230, 320, 325, 369–371,
velocity (ubiquitous, Ch13) 382, 384, 385, 386–388, 444, 536, 545
unconsolidated 4 Central 228
marine clastic sediments 430 continental 228, 249
sediments 7, 15, 205, 206, 243, 249 East Coast 254–256
sands 375–377 Eastern 228
undamaged rock 130 North Western 451
Index 713

Western 210–216, 221–224, 227, 228, 244, 245, independent- analyses in each azimuth 384
251–253, 513–515 interval stacking- 234
West Coast (Oregon, Washington, California) 231, inversion (sub-ocean floor) 281
257–259 lateral- variations 257, 258, 381
South-Western 210, 230 lithology 48
US Bureau of Mines 64 low- zones 71
US Department of Energy 382 low- layer 140, 141
US Geological Survey 216 model 93, 154, 157, 212, 219, 252, 253
Uzbekistan 211 model, multi-layer, sub-ocean 243, 252, 253, 258, 262,
266, 267, 269, 274, 289
Vacuum Field, New Mexico 451 model, multi-layer, altered zone 317
Vajon dam site 108 nine (or more) components of- 355, 361
Valhall reservoir 391, 392, 396, 438, 453–455 oscillation in well logs 299–301, 316
Vamanashi earthquake 213 permeability, of joints, of fractures 363, 364
Vardø sub-sea road tunnel, Norway 145 porosity-permeability 332–334
Vatnajökull eruption, Iceland 434 rock condition (Japan) 48
Veas sewage treatment caverns, Slemmestad, reduction in subsiding reservoir 399
Norway 148 residual- tomogram 57
velocity, velocities (see also numerous data sets at: VP) rosette 40
anisotropy 129, 299 thickness anomaly 256
anisotropy due to fractures 358–360 tomograms (see tomography)
anisotropy due to foliation 38, 248 undisturbed 313, 319
anisotropy due to joints and bedding 361–364 versus radius (see EDZ) 120, 318
anisotropy expressions 356 velocity-depth (see VP, by category)
anomalous, supposedly 271 curves 91, 92
anomaly 14, 256, 257 discussion 271, 272
azimuthal anisotropy 40–44, 129 gradients (see also velocity gradients)
contoured, continuous profile 284 non-uniqueness 271
contrast 15 structures, sub-continental crust 241, 242, 245–251
density space 22 structures, continental margins 254–261
depth models 369 structures, geothermal fields 394, 395
depth and crustal type, to 50 km 251 structures, mid-Atlantic ridge 261–273, 287–290
depth-age models 369–372 structures, East-Pacific rise 273–290
depth models 266, 267, 271, 274, 275, 369 velocity ratios (see VP and VS categories following)
discrepancy, lab to in situ 263, 265, 268, 276, 281 (Vo/VP)2 with azimuth (granites) 41
dispersion (see frequency dependence) 45, 386 (V/Vmax) with azimuth (limestones) 44
dispersion due to absorption 386 (V’/V) w.r.t. incident angle to fault 48
dispersion, negative 343, 344 (VF/VL) field to lab 69, 73
distribution 4 (VF/VL)2 field to lab squared 69, 74
false 146 Vema, North Atlantic fracture zone profile 273
fluctuation in interbedded strata 379 Venezuelan oil sand 330
gradient 4, 6, 52, 91–95, 265, 268, 271, 272, 273, 285, Ventura Basin 260
348, 370–371, 381 vertically-aligned fractures, jointing 374, 407,
gradient, extreme 14, 283, 286 409, 477
gradient, large, steep 15, 60, 222, 268, 269, 274, 277, causing anisotropy with low frequency 382
278, 348, 371 causing biased sampling with vertical wells 407
grid 16 causing signal distortion with mid-frequency 382
high- where Qseis and Qrock also high 410, 411 leading to reflections with high frequency 382
high- regions 129 missed by vertical wells 407, 408, 441
higher- with water than oil, in 4D 397 progression in complexity of matrices 416
hysteresis 31 vesicules, vesicular (see basalt)
increase with age 261–263, 276, 282, 287–290 gas-filled 265
increase due to compaction 398, 402 Vibrometric 56
714 Index

vibration VP – biaxial and axial stress (in situ jointed block of granite)
longitudinal 5 166
torsional 189 VP – biaxial and axial stress (in situ jointed block of sand-
vibrator (see source) stone) 167
Vicker’s microhardness 78 VP – confining pressure (diverse rocks) 31
Viking Graben, North Sea 352 VP – confining pressure: sandstones, dry and saturated 32,
viscosity bedded-coal, dry, saturated, 3-axis testing 201
visco-elastic 183, 475 VP – confining pressure and pore pressure (sandstones) 193
viscous VP – cross-hole (not tomography) 50–52, 66, 67, 106,
damping 190 119, 120, 134–136
dissipation 190 VP – cross-hole (time of heating and cooling) 135, 136
void (see joint void) VP – cross-hole tomography 54, 55, 58, 62–64, 132, 234
formation due to blasting 61 VP – cross-pillar tomography in mining 60, 61, 65
-space due to blasting 118 VP – density 19
volcanic VP – density (dolerite) 22
porous- rocks 27, 142, 251–253 VP – density (extreme confining pressure) 23
soil 25 VP – density (Tertiary foundations) 21
volume comparison VP – density (sulphide ores) 23
compaction-subsidence- with core plugs 400 VP – density (basalt, gabbro, serpentinite peridotite,
volumetric joint count 167 20 my) 265
velocity (see VP and VS , and all categories below) VP – depth (see multiple categories below)
azimuthally-dependent- 301 VP – Ee and Ed (pseudo-static: limestones and diverse rock
depth trends (see VP-depth, etc) 297 types 99–103
elliptical Vfast and Vslow replaced at faults 394 VP2 –Ee and Ed 100
fast- not equal matrix at faults 394 VP – Emass (pseudo-static) and other parameters, Tertiary
P-wave-, see VP specification below bridge foundations 21
radially-dependent- 302 VP – Edyn (diverse sources) 107
reduction, source of 306 VP – extensional strain amplitude: sandstone 187
zone of changed- 302 VP – failure stress (plate loading to failure) 104
VP (see also P-wave) alphabetic listing of categories VP – fractures per meter in drill holes 73
(see itemized data below alphabetic order) VP – joints per meter (diverse rocks, countries, weathering)
VP – age, sub-ocean, spreading ridge data 262, 264, 265, 72
287–289 VP – joint spacing, penetration strength, chalk 79
VP air-dry – VP saturated (dolerite) 23 VP – joint spacing, depth, shallow tunnel in chalk 79
VP anisotropy (parallel) and VP (perpendicular) slate 38 VP – load-unload (slate) 38
VP anisotropy (maximum) parallel to major joint set 40 VP – Lugeon value (permeability) 159, 160
VP – angle of incidence (anisotropic rocks, dry or saturated) VP – orthogonal directions (jointed blocks of granite: dry or
38 wet) 40
VP – artificially jointed: frequency, roughness, clay-filling, VP – penetration rate (TBM) 149
dilation/shearing 85, 86 VP – permeability (see listings below) 159–177,
VP – azimuthal 360° (in situ sparsely jointed granite) 40 333–337
VP – azimuthal 360° (in situ jointed granite limestones) 40, VP – permeability – FZI 333–335
42, 43 VP – porosity (see listings below) 19, 21, 24
VP – azimuthal 360° (in situ jointed) 44 VP – porosity (clean sandstones, and suspension) 324
VP – azimuthal 360° 3D VSP (0 to 520 m) 44 VP – porosity (crystalline, volcanic rocks) 28
VP – axial load (pre- and post-failure, coal) 104 VP – porosity – clay-content (sandstones) 25, 331–333
VP – axial stress (perpendicular, parallel: foliated VP – porosity (of rock mass, Tertiary foundations) (nr %)
gneiss) 39 19, 21, 22, 24
VP – axial stress (indentor, sandstone, tomography) 59 VP – Q-value, core logging, cavern site 76
VP – axial stress (basalt 126), (quartz-monzonite 136) VP – QP – depth plots (1800 to 2800 feet) 234
VP – axial stress (lab- and field-scale comparison, granite) VP – QP – depth trends (1 to 20 km) 230
166 VP – QP – depth trends (1 to 1200 km) 245
VP - axial stress-strain, loaded to failure (tuff-sandstone, VP – rock quality Q-value (see listings below) 13
granite) 195, 196 VP – rock quality – deformability 24
Index 715

VP – RQD – joints per meter, or core lengths (hard rocks, (iii) VP – depth (continental crust, shallow)
near-surface) 6, 13, 71, 72, 75 VP – depth – fractures/m (900 m XTLR well, crystalline
VP – rate of penetration, hardness, RQD, density 78 rocks near SAFZ) 90, 91
VP – reduction with freeze-thaw cycles 185 VP – depth (0 to 1.6 km, nuclear waste site tomography)
VP – saturation (foundation) 21 58
VP – Schmidt N-hammer rebound (crystalline, volcanic VP – depth (0 to 5.5 km, Lower Triassic sandstone in deep-
rocks) 28 est 1.5 km) 233
VP – seabed tomography 56 VP – VS – depth, on land (0 to 7 km, porous, jointed vol-
VP – support type 141, 147, 149 canics, tuff, NTS) 251–253
VP – temperature (ambient to freezing, sandstone) 33 VP – VS – depth – fractures/m – VP/VS ratio (1100 m
VP – three dimensional incident angles (truncated cube, MONT-1 well, jointed granodiorites) 90, 92
shale) 39 VP – VS – transmissivity, and VP/VS (400 to 1800 m deep
VP – time of drying out (granite) 31 boreholes in marl) 161
VP – time of heating and cooling 185
VP – uniaxial compressive strength (crystalline, volcanic (iv) VP – depth (continental margin)
rocks) 28 VP – depth (and density) (0 to 45 km, also laterally 240 km)
VP – uniaxial compressive strength (dolerite) 23 US East Coast 255, 256
VP – uniaxial compressive strength (diverse) 25 VP – depth (4.4 to 6.0 km) Hawaiian ridge 258
VP – uniaxial compressive strength (Tertiary mudstones, VP – depth (0 to 6 km), NE Atlantic, sub-basalt profile 403
sandstones) 27 VP – depth (0 to 30 km) San Francisco Bay area crust 259
VP – uniaxial compressive strength (shale) 26 (v) VP – depth (sub-ocean, spreading ridge)
VP – water level fluctuations (reservoir dam site) 89, 90 VP – depth East Pacific rise 273–287
VP – water saturation – porosity (crystalline, volcanic VP – depth WAP 10, 52 km, East Pacific rise 284–286
rocks) 29 VP – depth 14°S on East Pacific rise 286
(i) VP – depth, extreme depth VP – depth Mid-Atlantic ridge 243–245, 261–273
VP – VS – depth (0–6300 km) crust to inner core 241 VP – depth Ontong-Java 206
VP – VS – confining pressure (extreme, 5–10 kb) 32 VP – depth 0 to 1 km, East Pacific rise, linear gradient
assumption 275
(ii)VP – depth (continental crust, deep) VP – depth sonic log, Hole 504B, Costa Rica ridge 280
V P – depth (0 to 8 km, six geothermal fields, Kenya VP – depth – age young oceanic crust 276
Rift) 395 VP – depth – permeability – porosity, Hole 504B, Costa
V P – depth (0 to 9 km, KTB super-deep well, sonic and Rica E.Pacific ridge 278
VSP) 249, 250, 252 VP – depth – porosity – permeability (0 to 500 m of sedi-
VP – depth (5, 10, 15, 20, 25 km histograms) 247 ments) marine logging and modelling, Site 977, ODP
VP – depth (0–25 km, 0–50 km) 247, 248, 250, 251 Leg 161, Shipboard Scientific Party 336
VP – depth (0–35 km) and lateral variation (0–200 km) VP – VS –depth, VSP, Hole 504B, Costa Rica ridge 280
229 VP – VS –depth, VSP, Hole 504B, Costa Rica ridge, and lab
VP – depth (0 to 50 km) five types of continental crust, comparison 281
with lab-sample comparison 250 VP – VS –depth 264, 266, 267, 275, 277, 280, 281, 286
VP – depth (0 to 50 km) Proterozoic, Phanerozoic, VP – VS – Poisson’s ratio, young crust 270
Platform, Oregen crust 251
VP – depth (5, 10, 15, 20, 25 km) orogens, shields and plat- (vi) VP – depth (petroleum reservoirs, shallow, deeper)
forms, continental arcs, rifts, extended crust, average VP – depth, also QP /VP – depth, 210–400 m, sands and
crust 248 sand channels 375–377
VP – depth (0 to 50 km) continental crust, USA VP – depth (0 to 5500 feet) (also Q) Melville Island,
N.W.Montana to Washington State 251 Canadian Arctic 233
VP-VS – depth, on land (0 to 9 km, KTB super-deep well, VP – depth (0 to 6 km), NE Atlantic, sub-basalt
sonic and VSP) 249, 250, 252 profile 403
VP – VS – QP–QS - depth (0 to 35 km, and laterally 0 to VP – depth (0 to 2.3 km), limestones, clays, over-pressured,
190 km) Japan 229 two wells 450, 452
VP (variation) for 29 crustal rock types, 309°C, 20 km VP – depth (1800 to 2800 feet) (also Q) mixed lithology,
equivalent depth 249 BP Devine Test Site 234
VP (anisotropy) matrix, high pressure, high temperature VP – depth (4000 to 4350 feet) (also Q) fractured shales,
248 Buena Vista Hills 237
716 Index

VP – depth (0 to 5 km) high pressure well, over-pressure VP static/VP dyn (and Young’s modulus) – axial strain magni-
effects 298 tude, Boise sst. 340
VP – depth (0 to 5 km) and lateral, at salt dome, North Sea VP static/VP dyn (and Young’s modulus) – axial strain magni-
381 tude, Pecos sst. 340
VP – depth (850 to 1000 m) also core porosity-core perme- VP static/VP dyn (and Young’s modulus) – axial strain magni-
ability 334 tude, Ohio sst. 340
VP, VS – depth (0 to 1 km), CBTF well, also density, per- VP and VS – 3D confining pressure, before/after fracturing,
meability, birefringence 445 cubic specimens, Crossland Hill sst. 362
Vph and Vpv – confining and pore pressure (Berea,
(vii) VP – depth (near-surface, land-based)
Michigan sandstones) 192
VP – depth (0–50 m) and lateral variation, various sedi-
ments and rock 10 sandstone, tight gas
VP – depth slice (mean 32, 96, 106 m and laterally, allu- stress-strain data 339
vium, lithified sands-sandstone) 374, 375 velocity-density data 327
VP – depth (1–15 m, shallow soils) 24 VP –VS data 328
VP – depth (5–15 m, 6 m,120 m, 200 m, mostly rock) 79, 3D VP –confining pressure, spherical samples,
52, 52, 20 layering/jointing, dry, saturated 358
VP – depth (0–60 m approx., jointed gneiss, tomography) 3D permeability –confining pressure, spherical samples, lay-
55, 56, 88, 89 ering/jointing 359
VP – depth (mine-drift walls) 70, 73
unspecified sandstones
VP – depth (10–500 m approx., into tunnel walls: EDZ
VP – porosity, high porosity sandstones 325
related) 119–121, 123–125, 127–129, 133
VP – density, high porosity sandstones 327
VP – pressure/depth, core and sonic log comparison,
VP – density, sandstones 326, 327
Whitchester sst. 348
VP – porosity clay content % 331–333
VP – VS (0 to 35 m) (also dynamic elastic parameters) 111
VP – porosity-velocity, clean sst. and sand suspension 324
VP – VS (5 to 40 m) hard igneous, metamorphic rocks 6, 105
VP – porosity, sandstone and sand suspension 324
VP – VS – depth (0 to 60 m), Conoco borehole, fractured
VP – saturation % gas/brine, sandstones 341
limestones, impermeable shales 448
VP – saturation, patchiness data, multiple clean/shaly sand
VP – VS large number of rock types, near-surface refraction
composites 33, 337
seismic assumed 12
VP – frequency – 1000/Q fused glass beads 345
VP – VS from fine resolution of sonic logs, interbedded
VP – VS data, sandstones 328, 329, 372
shale, sandstone, limestone 379
VP – VP/VS – degree of saturation, sandstone 189
VP/VS – depth (0 to 70 m) clay over mudstones 8
VP and VS versus differential pressure 298
VP/VS ratio: near-surface 6, 9, 161, 252
VP and VS versus excess pore pressure 298
VP/VS ratios, six geothermal fields, 0 to 8 km, Kenya Rift 395
VP (and QP) – effective pressure, microcracked sandstone
VF/VL ratio of field (seismic) /laboratory (ultrasonic) 105,
347
106
VP/VS)2 – QP/QS – (and pore pressure) to distinguish sand
(viii) VP,VS rock physics data sets and sandstone 351
specified sandstones VP/VS – differential pressure, also to extremely low pressure,
VP – porosity, Troll sand 326 sands 351
VP (and QP) – frequency – effective pressure, dry or brine
oil sands/tar sands
saturated, Berea sst. 344
VP – temperature – oil/gas %, Kern River oil sand 330
VP (and QP) – frequency – effective pressure, dry or brine
VP – temperature – brine/oil %, Venezuelan oil sand 330
saturated, Boise sst. 344
VP – temperature, Canadian tar sand 331
VP (and QP) – frequency – effective pressure, dry or brine
saturated, Massilon sst. 344 diverse rock types
VP – VS – temperature, Boise sst. 331 VP – porosity, limestones, dolomites, chalks 325
VS – permeability, fracture closure, cubic specimens, VP – density, limestones, dolomites, anhydrite, rock salt,
Crossland Hill, Fontainebleau, Springwell sst. 364 shale 326
VP – VS – axial strain (Berea sandstone) 37 VP – density, limestones, dolomites, chalks 327
3D qP – normalized permeability, stereograms, North Sea VP – VS limestones, dolomites, chalks 328
sst. 390 VP – VS shales, mudrock 372
VP static/VP dyn and Young’s modulus versus axial strain mag- VP – VS – kerogen %, Kimmeridge shale, North Sea, 0° and
nitude, Berea sst. 340 90° 353
Index 717

VP – pressure, shale, 0°, 45 ° and 90° shear wave splitting with- focussed on domains 410, 439
VP – VS (and Vsh and Vsv) – pressure, shale, 0°, 45 ° and 90° surface shot points, single azimuth 382, 383
354 vertical-cable seabed- 446
VP – (and Vsh and Vsv) – 0° to 90° from bedding, and stere- walk-around- 383, 439, 440
ographic, shale, 354 walk-away- 382, 383, 439
walk-away- for S-wave acquisition 439
qP component P-wave velocities through truncated cube,
wide-aperture layout- 383
shale 39
zero offset- 383
VP – VS – axial stress (quartz-monzonite 136) vug, vugs 196, 197
VP – VS – confining pressure, Barré granite 36 exact structure 196
VP – VS – density (basalt, gabbro, serpentinite peridotite vuggy
20 my) 265 porosities 196
VP – VS – confining pressure, marble columns 81
VP – VS (dry, crystalline, volcanic rocks) 29 Wales 9, 11, 160
VP – VS – group velocity versus stress, mated, unmated ten- walkaway VSP (see VSP)
sion fractures 87 WAP 10 (52 km profile), East Pacific Rise, mid-ocean ridge
VP and VP/VS for dry, partly saturated, saturated: extension 283–285
tests 189 water 4, 12, 254.
bearing joint set 42
VP/VS diverse data
bearing sandstone reservoirs 343
VP/VS energy ratio (Es/Ep) 131
cold- injection, HDR 414
VP/VS ratio: theory 6, 7, 109
conducting joints, spacing with depth 488, 490
VP – rock quality
conducting joint/fracture/fault directions 384, 541–543
reversed- 17
conducting fracture directions under shear stress 402, 541,
VP – Q – depth – porosity (strength) relation 271, 273,
542
274, 285, 285
content 27, 169
VP – Q – M (deformation modulus) model 83, 92, 94,
depth 296
108, 132, 134
expulsion 135
VP – Q – M – L (Lugeon) model 164, 225, 291
fault-eroding water pressure 154, 156
VP – Q – M – permeability, near-surface model 225, 291
filled fissures 22
VP – Q – M – permeability, near-surface model, curve
filled structure 281
jumping 291
filled holes 61, 73
VP – Q or Qc model 6, 7, 13, 18, 47, 74, 115, 271
flooding (see water flooding)
VP – Qc – gradient, sub-ocean basalts 287
fresh- effect 297
(ix) rock physics/poro-elatic modelling ground- 11
VP – qP – VS – frequency, matrix and 10 cm fractures 466 ground-water pressure 159
VSP 14, 49, 207–209, 249, 369, 377, 380, 387 injection tests 162
azimuthal- 384, 385 injection into reservoir, effect on joints 487
diametrically-opposed- 445 inrush 151
down-hole receivers 382, 383 pressure anomaly, in well 433
effect of subsidence on- 400 pumping test (extraction) 162, 173, 174
far-offset 439 saturated joint samples 509, 510
-for horizontal well sections 382 saturation (see saturated) 7, 9, 27, 29, 200, 201, 352, 353
for QP and QS analysis 386 saturation weakening of chalk 399–402
for analysing velocity differences re sonic 386 sensitive reservoir rocks 402
limited offset- 383 table 8, 15, 124
multi-azimuth- 382, 383 velocity of- 159
multi-azimuth- for fracture anisotropy detection weakening modelling, chalk 399, 400, 402
382–386 water-flood, -flooding 329
multi-azimuth reversed- 385, 447 causing local cooling and aperture increase 329, 402
near-offset- 377, 384–386, 445, 446 causing contraction of matrix 402
nine-component- 385, 471–475 caused reduced velocity, amplitude, frequency 353
opposite-azimuth- for fracture dip 445 conjugate shearing 390, 522
-QP less than sonic-based QP 386 directionality 402
718 Index

water-flood, -flooding (contd) well (see also borehole)


directionality of flow assumed 402 better placement, deviation, with shear-wave technology
directionality of flow measured 403 407
faster break-through due to 402 data, typical VP , QP , n% 345
fracture and fault opening due to- 402 deviation 305
HTM effects during- 401 deviation for improved sampling of structure 407,
model of intermittent- 353 408
pressure-drive effect 264 diameter, largest 304
tracking- injection fronts 397 flow-rate versus P-wave anisotropy % 405
water injection 402 high flow-rate- 405
for compaction control, Ekofisk 399–402, 442 high pressure- 298
wave, waves horizontal section of-, collapse analogue 320
cut platform (photo) 488 location in reservoir 297
paths, horizontal 371 -logging developments, dipole 302, 313, 314
scattering 5, 372 -logging interpretation 302
transient 17 -logging while drilling, LWD 301, 302, 312–320
wave fronts 16 low flow-rate- 405
wave length rock/fluid pressure aspects 295–299
comparison, reflection, teleseismic 410 sections 369
dominant- long compared to 372 size 309
in relation to thin bed 372 Wendover, Utah 238
wave speed expressions 5, 477 Widemouth Bay, Bude, Devon, England 488
weak rock Wilmington Field, Long Beach, Caliornia 400, 454
at reservoir depth 534 WIPP, New Mexico 123
confined (␴1- ␴3) increase 534, 539 wireline (sonic) logging 302, 311, 313, 499, 500
crossing critical state line 534, 539 monopole and dipole logs 318, 319
improved JCS at reservoir depth 534 monopole and dipole log comparison with LWD 318
intolerance of stress difference 297–299, 302–304 temperature limits 320
limits of strength increase due to pore collapse 534, 539 velocities compared with LWD velocities 318
weakness zones 145 velocities, compressional and shear 318, 319
weathered 4 Wood’s metal 197
constituent minerals 24 world
deeply 4, 30 first TBM tunnel 319
extremely- 103, 114 map 246, 251
in place 4 map model tiles 550x550 km 249
seams 14 record speeds (TBM) 17
unweathered 13, 71, 74 supply of oil, gas, water (e⬍E detail) 488
zone 19, 20 -wide compilation 244–251
weathering 3, 26, 27, 324 Worthington, Lubbe, Hudson range of compliances 503,
caused by microcracks 225 504, 506
degree of-, joint/rock rebound ratio r5 /R5 510 Wulff ’s stereogram 39
gradational 56 Wyoming, USA 165
seafloor (sub-ocean) 281
zone of- 28 Xiaolangdi multipurpose dam, China 74
weight X-hole (see cross-hole) 208
weighted-average Qseis response X-ray
w.r.t . finely interbedded rocks 380 computerized tomography 196
wellbore (see borehole, well) diffraction technique 359, 360
alteration (see alteration) XTLR well, Mojave Desert 90, 91
‘ballooning’ (see HPHT) 320, 321
damage (see EDZ) 301, 317 Yangtze river, China 172
stability 339 yield pillar in coal 60, 61
Wellenberg, Switzerland 161 Yo-yo mid point 67
wellhead 43, 44 Young’s modulus (see also deformation, dynamic) 273
Index 719

axial dynamic modulus (␺) 5 ZEDEX (zone of excavation disturbance experiment)


E-modulus (from elastic pseudo-static stress/strain of intact 131, 132
specimen) 191, 198 zero
static (see deformation modulus) 339, 340 age crust, velocities 267, 276, 285
Edyn (from VP and VS) 13, 57, 75 -offset vertical seismic profiling 383, 434
Yuan-Lin site, Taiwan 205 ZN and ZT (see compliance)
Yucca Mountain borehole data 543 Zoback and colleagues stress measurements 429
Yugoslavia 49–51, 97, 98, 101, 102, 109, 171 Zoback and colleagues conducting-fracture measurements
541, 542
Zavoj dam site 171
Zechstein (salt dome) North Sea 381
721

Plate 1 Deep (1000–1200 m) cross-hole tomography at the UK Nirex Ltd Sellafield site. (Schlumberger GeoQuest, Nirex Report S/94/007,
by kind permission). (Figure 4.10).
722

Plate 2 An igneous intrusion (dike) tends to have elevated perme-


ability due to the number of joint sets (typically four: (a)
Jn  15). High-modulus mineralized veins may be frac-
tured by subsequent tectonic deformation, helping to
maintain some permeability despite the ‘sealing’ process.
(Figure 11.75).

(b)

Plate 3 Fracture development over time in a FRACOD model of a


circular opening in a jointed zone. Note the ‘rotation’ of the
diametral-pair of red regions, which represent low factors of
safety against shear failure. Further fracturing dissipates and
displaces the low F. of S. zones, suggesting that more fractur-
ing could occur across the ‘E-W’ diameter. Changed seismic
response over time is easy to imagine, also a mud-filtrate
invasion speed that could be highly non-uniform, due to
developing permeability in the partly connected disconti-
nuum. Nick Barton & Associates 2005 contract report.
See Figure 12.15 for input data. (Figure 12.16).
723

Plate 4 Comparison of a) detailed seismic-based (4D time-shift) compaction interpretation (with adjustment for the velocity reduction
caused by a subsiding overburden), with b) geomechanics-based one-dimensional strain compaction model, that included porosity
reduction due to weakening effect of water saturation. Smith et al., 2002. Note gas cloud effect in centre of seismic model.
(Reproduced by kind permission of Norwegian Petroleum Society, NPF). (Figure 14.29).

(a) (b)

Plate 5 Two examples of attenuation modelling with a set of vertical aligned fractures, using Chapman’s dynamic poroelastic matrix-and-
fracture-set model. See Chapter 15 for a description of this model. Maultzsch et al., 2005. (Reproduced by kind permission,
Maultzsch pers. comm. 2005). (Figure 14.37).
724

(a) (a)

(b)
(b)
Plate 7 Bluebell-Altamont Field anisotropy interpretation, from
Plate 6 Bluebell-Altamont Field anisotropy interpretation, from
Liu et al., 2003a, with kind provision of files from Liu,
Liu et al., 2003a, with kind provision of files from Liu,
pers.com. 2005. a) Steep time delay gradient in reservoir
pers.com. 2005. a) Except at low frequency (0–10 Hz), a
interval, with frequency dependence. b) Interpretation
reasonably constant polarization at 40° to 45° is shown. b)
of anisotropy percentage as function of frequency.
Time delays show three intervals: gradient, flat, gradient,
(Figure 15.57).
implying anisotropy, isotropy, anisotropy. (Figure 15.56).
725

(a)

(b)

Plate 8 a) The relative error between the predicted and measured


time-delay/depth, evaluated over four frequency values for
a range of possible fracture densities and sizes, for compar-
ing with multi-component shear wave VSP data acquired
in the Bluebell-Altamont field in Utah. b) The rms error
zoomed around the minimum, where the error is less than
5%. (Chapman model application, by Maultzsch et al.,
2003 and Liu et al., 2003b). (Figure 15.58).

(c)

Plate 9 A wave-cut platform in a jointed dolomite bed. These beds


occur at intervals in the Kimmeridge shale, outcropping in
Kimmeridge Bay, Dorset, England. The joints show a)
implied JCS  c due to weathering and preferential
wave erosion, and b) implied (local) JCS  c, due to
subsequent mineralization of dominant conducting joints.
c) A fine example of joint cementation which may prevent
the use of normal joint characterization techniques.
Widemouth Bay, near Bude, Devon, England. (Figure 16.5).
726

Plate 10 The photograph is a fine example of contrasting JRC values from an interlocked joint and a minor fault in welded tuff, with respective
JRCo values of about 15 and 1. (Figure 16.28).

Plate 11 Hydraulic aperture (e), versus normal stress ( n), versus average rock temperature (T° C) in the permeability test volume of the
TerraTek heated block test, CSM mine, Colorado. Note aperture (e) reductions from 30.0
m to 18.3
m, to 12.9
m and finally
to 9.1
m as a result of temperature rise, despite constant applied stress. This gives an apparent reduction in the normal stiffness
in this test, but in the warmth of a deep petroleum reservoir, would have allowed joints to remain stiffer since their formation.
Barton, 1982. (Figure 16.33).
727

Plate 12 Interlocked and sheared joints in ‘wavy’ columnar basalt, demonstrating the role of asperities and dilation on aperture distribution.
Columbia River Basalts, Washington State, USA. (Figure 16.37).

Plate 13 Vertical view through the inter-bedded bituminous shales and dolomite bed, in Kimmeridge Bay, showing a) complex, b) ordered
sub-vertical joint patterns. Dorset coast, S. England. (Figure 16.39).
728

Plate 14 Normal and shear stresses for fractures identified as hydraulically conducting (closed symbols) or non-conducting (open symbols).
Cajon Pass (triangles), Long Valley (circles), Nevada Test Site (squares). Townend and Zoback, 2000, with data also from Colleen
Barton et al., 1995. (Zoback, 2006 pers. comm., by kind permission). (Figure 16.63).

Plate 15 Normal and shear stresses for fractures identified as hydraulically conducting or non-conducting, using borehole imaging. Cajon
Pass (red diamonds and dots), Nevada Test Site (green circles and dots), Long Valley (yellow triangles and dots), KTB (Germany –
blue squares and dots). Inset shows
 / n for combined data set. Zoback and Townend, 2001, with data from Ito and Zoback,
2000, and from Colleen Barton et al., 1995. (Zoback, 2006 pers. comm., by kind permission). (Figure 16.64).
729

Plate 16 a) An example of a massive rock mass with rock quality Q  1000 and a deformation modulus in excess of 100 GPa. A very low atten-
uation is implied. b) The fault-collapse blocking the tunnel on the right would give almost the lowest rock quality Q  0.001, and a
modulus of deformation lower than 1 GPa. It is perhaps ‘off-the-scale’ regarding the conventional definition of Qseis, and would need
to be under stress to allow spectral analysis of measurable amplitudes. Its Q-value would then be higher too. (Figure 16.79).

Plate 17 The reality of near-surface construction of tunnels and caverns in rock. Note the three joint sets causing deep over-break. (Figure B1).

S-ar putea să vă placă și