Sunteți pe pagina 1din 178

EVALUATION OF SHANSEP PARAMETERS AND INITIAL-SHEAR-STRESS INDUCED

STRENGTH ANISOTROPY OF A LAGOONAL DEPOSIT NEAR KE’EHI INTERCHANGE,


HONOLULU, HAWAI’I

A THESIS SUBMITTED TO THE GRADUATE DIVISION OF THE UNIVERSITY OF HAWAI’I AT


MĀNOA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF

MASTER OF SCIENCE

IN

CIVIL ENGINEERING

DECEMBER 2014

By

Todd K. Kawamoto

Thesis Committee:
Phillip Ooi, Chairperson
Horst G. Brandes
Peter G. Nicholson
ABSTRACT
Four CKoUC, four CKoUE and four CKoUDSS tests were conducted to study the
effects of initial- shear-stress-induced anisotropy on the strength of a Hawai’i lagoonal
deposit classified as an elastic silt (MH). Undrained strength ratios were plotted versus
OCR, and Stress History and Normalized Soil Engineering Properties (SHANSEP)
parameters were determined. The SHANSEP parameters were then used to calculate
undrained strengths based on the stress history of the lagoonal deposit. These values
were then compared to undrained strengths from 11 unconsolidated undrained (UU)
triaxial tests.

If the undrained shear strength, Su is assumed to be half the maximum deviator


stress (q) for the triaxial tests, the CKoUC strengths were largest, the CKoUE lowest and
the CKoUDSS in between. If Su = qcosϕ’ (where ϕ’ = effective friction angle), the CKoUC
and CKoUE curves shift downwards, resulting in the CKoUDSS curve almost coinciding
with the CKoUC curve. The effective friction angles ϕ’ were approximately 40o and 26o
for the normally and overconsolidated samples, respectively. The SHANSEP parameter S
for the Hawai’i lagoonal deposit was the largest compared to soils from other
publications.

An acid test to detect the presence of calcium carbonate in the lagoonal deposit
was performed. Vigorous bubbling and effervescence indicated the presence of
calcareous fines in the lagoonal deposit. This provides a possible explanation for the
large normally consolidated φ’ as well as the large SHANSEP parameter S.

Using CKoUDSS SHANSEP parameters, calculated undrained shear strengths were


within 10% on average of those from UU tests. In addition, the SHANSEP undrained
strengths were slightly more conservative than the 11 UU strengths in general.

i
ACKNOWLEDGEMENTS
The author would like to first thank the state of Hawai’i Department of
Transportation for their financial support. The author would also like to thank Geolabs,
Inc. for sampling, drilling and performing the CPT soundings. The author would like to
thank Mr. Jason Seidman for taking the time to show the author how to set up and run
the direct simple shear tests. Mr. Jose Bumatay and Mr. Landon Kaya assisted with
some of the index testing and consolidation testing, for which the author is very
thankful.

The author would like to acknowledge Mr. Reza Rahimnejad for sharing his data,
knowledge and lighthearted attitude throughout the years together, and wishes him the
best on his doctorate completion and future endeavors. The author would also like to
acknowledge his parents for their overwhelming love and support throughout the
Masters program. The author would like to acknowledge the thesis committee
consisting of Professors Phillip Ooi, Horst Brandes, and Peter Nicholson for reviewing
this work. Most of all, the author expresses eternal gratitude to Professor Ooi for his
invaluable knowledge and advice, utmost patience, and incredible support.

ii
Table of Contents

ABSTRACT.......................................................................................................................................... i
ACKNOWLEDGEMENTS .................................................................................................................... ii
Table of Contents ............................................................................................................................ iii
List of Figures .................................................................................................................................. vi
List of Tables ...................................................................................................................................xii
1 INTRODUCTION ........................................................................................................................ 1
1.1 Background ...................................................................................................................... 1
1.2 Objectives ........................................................................................................................ 1
1.3 Organization of Thesis...................................................................................................... 2
2 LITERATURE REVIEW ................................................................................................................ 4
2.1 Soil Anisotropy ................................................................................................................. 4
2.2 Effect of Intermediate Principal Stress ............................................................................ 4
2.3 Practical Relevance of Laboratory Strength Testing ........................................................ 5
2.4 Strain Compatibility ......................................................................................................... 8
2.5 Stress History And Normalized Soil Engineering Properties (SHANSEP).......................... 9
2.6 Definition of Undrained Shear Strength ........................................................................ 11
2.7 Previously Measured Undrained Strength Ratios and SHANSEP Research ................... 12
2.7.1 Triaxial Tests........................................................................................................... 12
2.7.2 Direct Simple Shear Tests....................................................................................... 17
2.7.3 Combined CKoUC, CKoUE and CKoUDSS Data ......................................................... 25
3 SITE CHARACTERISTICS........................................................................................................... 27
3.1 Site Location................................................................................................................... 27
3.2 Site Geology ................................................................................................................... 28
3.3 Soil Stratigraphy ............................................................................................................. 32
3.4 Test for the Presence of Calcium Carbonate ................................................................. 32
3.5 Soil Characteristics ......................................................................................................... 33
3.5.1 Index Properties ..................................................................................................... 34
3.5.2 Preconsolidation Pressure ..................................................................................... 37
4 TRIAXIAL AND DSS TESTING - SPECIMEN PREPARATION ....................................................... 41
4.1 Soil Extrusion.................................................................................................................. 41

iii
4.2 Triaxial Specimen ........................................................................................................... 42
4.2.1 Trimming and Dimensions ..................................................................................... 42
4.2.2 Filter Strips ............................................................................................................. 42
4.2.3 Drainage ................................................................................................................. 43
4.2.4 Rubber Membrane ................................................................................................. 44
4.2.5 Cell Chamber Apparatus ........................................................................................ 44
4.3 Direct Simple Shear Specimen ....................................................................................... 46
4.3.1 Trimming and Dimensions ..................................................................................... 46
4.3.2 Wire-Reinforced Rubber Membrane ..................................................................... 47
4.3.3 Sample Assembly ................................................................................................... 48
5 TRIAXIAL AND DSS TESTING – EQUIPMENT, SOFTWARE AND TEST PROCEDURES ............... 49
5.1 Triaxial Testing ............................................................................................................... 49
5.1.1 Equipment .............................................................................................................. 49
5.1.2 Apparatus ............................................................................................................... 52
5.1.3 Calibration .............................................................................................................. 53
5.1.4 Testing Phases and Software Input ........................................................................ 54
5.1.5 Corrections ............................................................................................................. 55
5.2 DSS Testing Program ...................................................................................................... 60
5.2.1 Equipment .............................................................................................................. 60
5.2.2 Apparatus ............................................................................................................... 64
5.2.3 Calibration .............................................................................................................. 66
5.2.4 Testing Phases and Software Input ........................................................................ 67
5.2.5 Corrections ............................................................................................................. 68
6 RESULTS ................................................................................................................................. 71
6.1 Triaxial (CKoUC and CKoUE) ............................................................................................ 71
6.1.1 Deformed Sample Shapes During Shear ................................................................ 73
6.1.2 Maximum Obliquity ............................................................................................... 74
6.1.3 Effective Stress Paths and Determining φ’ ............................................................ 75
6.1.4 Skempton’s Pore Pressure Parameter ................................................................... 78
6.2 CKoUDSS ......................................................................................................................... 80
6.2.1 Pore Pressure ......................................................................................................... 81
6.3 SHANSEP Curves............................................................................................................. 82
6.4 SHANSEP Su vs Su from Unconsolidated Undrained Triaxial Tests ................................. 85

iv
7 SUMMARY AND CONCLUSIONS ............................................................................................. 88
7.1 Summary ........................................................................................................................ 88
7.2 Conclusions .................................................................................................................... 88
7.3 Recommendations for Future Research ........................................................................ 90
References ..................................................................................................................................... 91
Appendix A: Boring Logs (Geolabs, Inc, 2011) .............................................................................. 96
Appendix B: CPT Logs (Geolabs, Inc, 2011) ................................................................................. 110
Appendix C: Detailed Testing Procedures for Triaxial Tests (Triaxial, 2005) .............................. 113
Appendix D: Detailed Testing Procedures for DSS Tests (Geonor, 1999) ............................... 121
Appendix E: Consolidation Stress-Strain Curves ......................................................................... 145
Appendix F: UU Stress-Strain Curves .......................................................................................... 158

v
List of Figures

Figure 1: Stress Systems Achievable by Various Shear Devices for CKoU Testing (modified
from Germaine 1982), from Ladd (1991) as published in Ladd and DeGroot (2003) ........ 5
Figure 2: Relevance of Strength Tests to Various Field Practices, from Kulhawy and
Mayne (1990) ...................................................................................................................... 6
Figure 3: Comparison of Effective Friction Angles Between Isotropic and Anisotropic
Undrained Compression Tests on Normally Consolidated Soils, data from Mayne (1985)
and Nakase and Kamei (1988), as published in Kulhawy and Mayne (1990) ..................... 7
Figure 4: Comparison of Undrained Strength Ratios Between Isotropic and Anisotropic
Undrained Compression Tests on Normally Consolidated Soils, data from Mayne (1985),
Nakase and Kamei (1988), and Mayne and Stewart (1988), as published in Kulhawy and
Mayne (1990) ...................................................................................................................... 8
Figure 5: Example of the Strain Compatibility Technique (after Koutsoftas and Ladd
1985), from Ladd (1991) as published in Ladd and DeGroot (2003) .................................. 9
Figure 6: Direct Simple Shear Apparatus at Norwegian Geotechnical Institute, from
Bjerrum and Landva (1966) .............................................................................................. 19
Figure 7: SHANSEP Curves of Five Different Normally Consolidated Clays, from Ladd and
Edgers (1972) .................................................................................................................... 20
Figure 8: Histogram of Normalized Undrained Strength for 100 Normally Consolidated
CKoUDSS Tests, from DeGroot et al. (1992) ...................................................................... 21
Figure 9: Histogram of Shear Strain at Maximum Shear Stress for 100 Normally
Consolidated CKoUDSS Tests, from DeGroot et al. (1992)................................................ 21
Figure 10: Shear Strain at Maximum Shear Stress vs. Plasticity Index for Normally
Consolidated CKoUDSS Tests on Cohesive Soil, from DeGroot et al. (1992) .................... 22
Figure 11: Normalized Undrained Strength vs. Shear Strain at Maximum Shear Stress
for Normally Consolidated CKoUDSS Tests on Cohesive Soil, from DeGroot et al. (1992) 23

vi
Figure 12: Comparison of Field Vane and Laboratory Undrained Strength Ratios for
Normally Consolidated Non-Varved Sedimentary Soils, from Ladd (1991), as published in
DeGroot et al. (1992) ........................................................................................................ 24
Figure 13: OCR vs. Undrained Strength Ratio and Shear Strain at Failure for Ko
Consolidated Undrained Tests: (a) AGS Plastic Marine Clay (PI = 43%, LI = 0.6) via
SHANSEP (Koutsoftas and Ladd 1985); and (b) James Bay Sensitive Marine Clay (PI =
13%, LI = 1.9) via Recompression (B-6 data from Lefebvre et al. 1983) [after Ladd 1991],
as published in Ladd and DeGroot (2003) ........................................................................ 25
Figure 14: SHANSEP Curves for Various Shear Modes of Singapore Residual Soils, from
Meng and Chu (2011) ....................................................................................................... 26
Figure 15: Peninsula Site Location on O’ahu, Hawai’i, images taken from Google Earth
(2013) ................................................................................................................................ 27
Figure 16: Approximate Boring Locations, image taken from Google Earth (2013) ....... 28
Figure 17: Excess Hydrostatic Head with Depth at Ke’ehi Interchange, from Clemente
(1984) ................................................................................................................................ 30
Figure 18: 1955 Map Showing the Estimated Peninsula Site Location Near Ke’ehi
Interchange, from Cline (1955) ......................................................................................... 31
Figure 19: Map Showing the Peninsula Site Near Ke’ehi Interchange, from Visher and
Mink (1964) ....................................................................................................................... 31
Figure 20: Cross-Section View of Groundwater Occurrence in the Southern Half of
O’ahu, Hawai’i. Diagram Courtesy of Honolulu Board of Water Supply, image taken from
Seu (2002): http://legacy.earlham.edu/~seuka/Hydrogeology_page/ ............................ 32
Figure 21: Test for the Presence of Calcium Carbonate on (a) the Hawai’i Lagoonal
Deposit and (b) a Tropical Residual Soil............................................................................ 33
Figure 22: Representative Grain Size Distribution Plots for Samples from Borings 4A, 5A
and 5C ............................................................................................................................... 35
Figure 23: Plasticity Chart for All Triaxial and DSS Samples............................................. 36
Figure 24: Atterberg Limits and Moisture Contents vs. Depth for Borings 4A, 5A and 5C
........................................................................................................................................... 37

vii
Figure 25: Sample Strain vs. Consolidation Pressure Plots for Samples from Borings 4A,
5A and 5C .......................................................................................................................... 38
Figure 26: Preconsolidation Pressure and Effective Overburden Stress vs. Depth for
Boring 4A ........................................................................................................................... 39
Figure 27: Preconsolidation Pressure and Effective Overburden Stress vs. Depth for
Borings 5A and 5C ............................................................................................................. 40
Figure 28: Soil Extruder and Extruded Sample ................................................................ 41
Figure 29: Trimming the Sample in the Trimming Apparatus ......................................... 42
Figure 30: Filter Strip Placement for Compression (left photo) and Extension (right
photo) Tests ...................................................................................................................... 43
Figure 31: Membrane Expander and Membrane ............................................................ 44
Figure 32: Complete Cell Chamber with Sample ............................................................. 46
Figure 33: Sample Trimming Apparatus .......................................................................... 47
Figure 34: (a) Wire Reinforced Membrane for DSS Tests and (b) Membrane Expander 47
Figure 35: Sample Assembly ............................................................................................ 48
Figure 36: S-type Load Cell for Triaxial Tests ................................................................... 49
Figure 37: Embedded LVDT Transducer for Vertical Displacements ............................... 50
Figure 38: LoadTracII Machine .......................................................................................... 50
Figure 39: FlowTracII Machine ......................................................................................... 52
Figure 40: Triaxial Apparatus ........................................................................................... 53
Figure 41: Extension Attachments ................................................................................... 53
Figure 42: Setup for Determining Young’s Modulus of the Rubber Membrane ............. 59
Figure 43: Comparator for Direct Simple Shear Tests ..................................................... 62
Figure 44: Shear Box ........................................................................................................ 62
Figure 45: Burettes for Specimen Saturation .................................................................. 63
Figure 46: Carbon Dioxide Tank for Specimen Saturation ............................................... 64
Figure 47: Direct Simple Shear Apparatus ....................................................................... 65
Figure 48: Specimen in the Direct Simple Shear Apparatus ............................................ 66
Figure 49: Monogram Digital Panel Indicator .................................................................. 67

viii
Figure 50: Load vs Time Plots for Direct Simple Shear Tests Without a Specimen ......... 69
Figure 51: q/σ’vc vs Axial Strain Plots for CKoUC Tests ..................................................... 71
Figure 52: q/σ’vc vs Axial Strain Plots for CKoUE Tests ..................................................... 72
Figure 53: Deformation Behavior for CKoUC (left photo) and CKoUE Tests (right photo) 74
Figure 54: Normalized ’1/’3 vs Axial Strain Plots for CKoUC Tests ................................ 75
Figure 55: Effective Stress Paths and Failure Envelopes for All Triaxial Tests ................. 77
Figure 56: Pore Pressure Parameter A vs Axial Strain Plots for CKoUC Tests .................. 79
Figure 57: Pore Pressure Parameter A vs Axial Strain Plots for CKoUE Tests .................. 79
Figure 58: Pore Pressure Parameter at Failure (Af) vs OCR for the Hawai’i Lagoonal
Deposit and Other Clays (Superimposed with Oslo, Oslo and London and Weald clays
after Simons, 1960 and Boston Blue clay after de La Beaumelle, 1991) .......................... 80
Figure 59: Normalized Shear Stress vs Shear Strain Plots for All Direct Simple Shear Tests
........................................................................................................................................... 81
Figure 60: Normalized Pore Pressure vs Shear Strain for All Direct Simple Shear Tests . 82
Figure 61: SHANSEP Curves for CKoUC, CKoUE and CKoUDSS Tests with Su = Maximum q
for Triaxial and Su = Maximum τh for CKoUDSS ................................................................. 83
Figure 62: SHANSEP Curves for CKoUC, CKoUE and CKoUDSS Tests with Su = q cos ’ for
Triaxial and Su = τh for CKoUDSS ........................................................................................ 84
Figure 63: Su Calculated from SHANSEP vs Su Obtained from UU Tests .......................... 87
Figure 64: Membrane in Expander ................................................................................ 131
Figure 65: Placing Cutter ................................................................................................ 131
Figure 66: Sample Trimming .......................................................................................... 132
Figure 67: Attaching Mounting Ring .............................................................................. 132
Figure 68: Placing Lower Filter Holder ........................................................................... 133
Figure 69: Locking Lower Filter Holder in the Mounting Ring ....................................... 133
Figure 70: Inverting Setup .............................................................................................. 134
Figure 71: Removing Disk ............................................................................................... 134
Figure 72: Mounting Caliper .......................................................................................... 135
Figure 73: Lowering Cutter Yolk ..................................................................................... 135

ix
Figure 74: Mounting Membrane and Expander ............................................................ 136
Figure 75: Mounting Top Yolk ........................................................................................ 136
Figure 76: Attaching Upper Filter Holder with Top Yolk ................................................ 137
Figure 77: Lowering Upper Filter Holder On Sample ..................................................... 137
Figure 78: Lowering Cutter Yolk ..................................................................................... 138
Figure 79: Installing Membrane ..................................................................................... 138
Figure 80: Membrane in Place ....................................................................................... 139
Figure 81: Placing O-rings .............................................................................................. 139
Figure 82: Weighing Assembly ....................................................................................... 140
Figure 83: Attaching Drainage Lines .............................................................................. 140
Figure 84: Securing Assembly on Base........................................................................... 141
Figure 85: Measuring 49 mm Space to Center Sample.................................................. 141
Figure 86: Lowering Vertical Ram to Barely Touch Sample ........................................... 142
Figure 87: Attaching Vertical LVDT ................................................................................ 142
Figure 88: Tightening Knobs for Shear Clamps .............................................................. 143
Figure 89: Shear Clamps in Place ................................................................................... 143
Figure 90: Burettes for Saturation ................................................................................. 144
Figure 91: Consolidation Stress-Strain Curve for Boring 4A SH-3 .................................. 146
Figure 92: Consolidation Stress-Strain Curve for Boring 4A SH-5 .................................. 147
Figure 93: Consolidation Stress-Strain Curve for Boring 4A SH-7 .................................. 148
Figure 94: Consolidation Stress-Strain Curve for Boring 4A SH-9 .................................. 149
Figure 95: Consolidation Stress-Strain Curve for Boring 4A SH-11 ................................ 150
Figure 96: Consolidation Stress-Strain Curve for Boring 4A SH-13 ................................ 151
Figure 97: Consolidation Stress-Strain Curve for Boring 4B SH-2 .................................. 152
Figure 98: Consolidation Stress-Strain Curve for Boring 4B SH-5 .................................. 153
Figure 99: Consolidation Stress-Strain Curve for Boring 4B SH-7 .................................. 154
Figure 100: Consolidation Stress-Strain Curve for Boring 4B SH-9 ................................ 155
Figure 101: Consolidation Stress-Strain Curve for Boring 4B SH-11 .............................. 156
Figure 102: Consolidation Stress-Strain Curve for Boring 4B SH-13 .............................. 157

x
Figure 103: Stress-Strain Curve for B-4A SH-3 UU Test ................................................. 159
Figure 104: Stress-Strain Curve for B-4A SH-5 UU Test ................................................. 159
Figure 105: Stress-Strain Curve for B-4A SH-7 UU Test ................................................. 160
Figure 106: Stress-Strain Curve for B-4A SH-9 UU Test ................................................. 160
Figure 107: Stress-Strain Curve for B-4A SH-11 UU Test ............................................... 161
Figure 108: Stress-Strain Curve for B-4A SH-13 UU Test ............................................... 161
Figure 109: Stress-Strain Curve for B-4B SH-2 UU Test ................................................. 162
Figure 110: Stress-Strain Curve for B-4B SH-5 UU Test ................................................. 162
Figure 111: Stress-Strain Curve for B-4B SH-7 UU Test ................................................. 163
Figure 112: Stress-Strain Curve for B-4B SH-9 UU Test ................................................. 163
Figure 113: Stress-Strain Curve for B-4B SH-11 UU Test ............................................... 164
Figure 114: Stress-Strain Curve for B-4B SH-13 UU Test ............................................... 164

xi
List of Tables

Table 1: Relative Effective Friction Angles of Normally Consolidated Tests, from Kulhawy
and Mayne (1990) ............................................................................................................... 7
Table 2: Undrained Strength Ratios from Five Different Normally Consolidated Clays, as
published in Ladd et al. (1977).......................................................................................... 13
Table 3: SHANSEP Parameter Results for Triaxial Compression of Boston Blue Clay, from
de La Beaumelle (1991) .................................................................................................... 13
Table 4: SHANSEP Parameter Results for Triaxial Extension of Boston Blue Clay, from de
La Beaumelle (1991) ......................................................................................................... 14
Table 5: Results from Ko-Consolidated Undrained Compression Tests of Resedimented
Normally Consolidated Boston Blue Clay at Various Strain Rates, from Sheahan et al.
(1996) ................................................................................................................................ 15
Table 6 Results from Ko-Consolidated Undrained Compression Tests of Resedimented
Overconsolidated Boston Blue Clay at Various Strain Rates, from Sheahan et al. (1996) 15
Table 7: CKoUC Results of Normally Consolidated Resedimented Boston Blue Clay at
Different Consolidation Stresses, from Abdulhadi (2012) ................................................ 16
Table 8: CKoUC Results of Overconsolidated Resedimented Boston Blue Clay at Different
Consolidation Stresses, from Abdulhadi (2012) ............................................................... 16
Table 9: Results of Constant Volume Direct Simple Shear Tests, from Bjerrum and
Landva (1966).................................................................................................................... 20
Table 10: Depths, Index Properties and Shear Tests Ran for Various Shelby Tube
Samples ............................................................................................................................. 36
Table 11: Specific Gravity Data and Sample Information for Borings 4A, 5A and 5C ...... 37
Table 12: Peak Strength Data, Maximum Obliquity Data and Sample Parameters for All
Triaxial Tests...................................................................................................................... 73
Table 13: Maximum Shear Stress Data and Sample Parameters .................................... 81
Table 14: Comparison of S and m values for the Hawai’i lagoonal deposit with those of
other soils.......................................................................................................................... 85

xii
Table 15: Comparison Between Su Calculated Using the CKoUDSS SHANSEP Equation
and Su Obtained from Unconsolidated Undrained Triaxial Tests ..................................... 86

xiii
1 INTRODUCTION
This chapter provides background information on the thesis work, explains the
objectives of this thesis and elucidates on the thesis layout.

1.1 Background
Soil strength is an important parameter needed in analyzing the stability of
foundations, retaining walls, slopes and other types of earth structures. Soil strength
interpretation is a critical step in stability analyses and should be performed with a
thorough understanding of the various factors that can affect soil strength. They
include strain rate, sample disturbance and anisotropy.

From a theoretical standpoint, soil properties are directionally dependent, and


ideally, engineers performing stability analyses should consider the effects of
anisotropy. However, accounting for strength anisotropy requires high quality sampling
along with complex soil strength testing. The equipment required for such tests are
cost-prohibitive and hence, such tests are not routinely performed in geotechnical
engineering practice in Hawai’i. Instead, most geotechnical engineers rely on
unconsolidated undrained triaxial testing to obtain the undrained shear strength of
cohesive soil at best. Sometimes, unconfined compression tests or correlations with SPT
blow counts are the norm. While many types of tests are available to study anisotropy,
only anisotropically consolidated triaxial compression, triaxial extension and direct
simple shear tests are conducted as part of this work.

While strain rate and sample disturbance are equally important, the scope in this
thesis work is focused on understanding the influence of initial shear stress anisotropy
on the undrained shear strength of a lagoonal soil in Hawai’i.

1.2 Objectives
The primary objectives of this thesis work include:

1. Conduct anisotropically consolidated undrained triaxial compression, triaxial


extension and direct simple shear tests on undisturbed samples of a lagoonal

1
deposit consolidated to beyond the preconsolidation pressure and then
unloaded to various overconsolidation ratios (OCR).
2. Apply the SHANSEP (Stress History and Normalized Soil Engineering Properties –
described in Section 2.5) technique to all tests to investigate how the normalized
undrained strength ratios vary with OCR and with the different types of shear
tests.
3. Derive SHANSEP parameters for the various types of tests and compare with
values published in the literature for other soils.
4. Compare the undrained shear strengths of this lagoonal soil from unconsolidated
undrained triaxial tests with those derived from the SHANSEP parameters and
the soil stress history.

1.3 Organization of Thesis


This thesis consists of 7 chapters.

Chapter 2 presents a literature review on soil anisotropy, effects of the


intermediate principal stress on shear strength, relevance of the various types of
laboratory tests to study soil strength anisotropy, strain compatibility considerations in
stability analysis considering soil strength anisotropy, SHANSEP technique, undrained
shear strength definition and some published undrained strength ratio data.

Chapter 3 presents information about the site where the lagoonal deposit was
sampled from. This includes the site location, geology, stratigraphy, results of chemical
analysis to test for the presence of calcium carbonate, soil index properties and soil
stress history.

Chapter 4 describes the soil specimen preparation procedures performed before


placing them into the testing devices. Setup procedures involve soil extrusion from
Shelby tubes and specimen preparation for triaxial and direct simple shear tests.

Chapter 5 describes the test equipment, software and test procedures. Details
on the necessary corrections are also provided for both triaxial and direct simple shear
tests.
2
Chapter 6 presents test results and discussion of the soil behavior for all tests.
This includes stress-strain behavior, sample deformation behavior, pore pressure data,
effective stress paths and undrained strength ratios. This chapter also provides plots of
undrained strength ratio vs overconsolidation ratio for the various test types and the
derived parameters that characterize soil strength anisotropy. Lastly, it compares the
undrained shear strengths of a lagoonal soil derived from unconsolidated undrained
triaxial tests with those derived from SHANSEP.

Chapter 7 closes the thesis work with summaries, conclusions and


recommendations for future research.

3
2 LITERATURE REVIEW
Described below are details relevant to this thesis work.

2.1 Soil Anisotropy


Soil anisotropy is defined as when a soil’s in situ property, such as strength,
stiffness, permeability, etc., is directionally dependent. The focus of this thesis is on
strength anisotropy.

The following paragraph describes five different types of soil anisotropy (Ladd,
1996). For natural clays with one-dimensional strain histories, inherent anisotropy
develops due to the soil’s depositional and consolidation histories, which can influence
both the soil’s preferred structure (or fabric) and material (e.g.; varved clays, fissures,
etc.). Initial shear stress anisotropy exists when the soil’s at rest lateral earth pressure
coefficient Ko is not equal to unity. This is also known as stress-system-induced
anisotropy. A third form of anisotropy involves a combination of inherent and initial
shear stress anisotropy. If an isotropic soil is pre-strained, it is said to have a stress-
induced type of anisotropy. The last type is termed evolving anisotropy, which occurs
when principal stresses rotate in the soil.

2.2 Effect of Intermediate Principal Stress


Stress system is defined as the direction of the major principal stress (σ1) with
respect to the vertical. Ladd and DeGroot (2003) explain that the effect of stress-
system-induced anisotropy can be illustrated using two variables: the intermediate
principal stress ratio, b, defined by:

𝜎2 − 𝜎3
𝑏= Eq. 1
𝜎1 − 𝜎3
where σ1, σ2 and σ3 = major, intermediate and minor principal stresses, respectively

and δ = direction of applied major principal stress at failure relative to the vertical
(depositional) direction (Ladd and DeGroot, 2003). These two variables (b and δ)
represent the effects of intermediate principal stress and stress-system anisotropy,

4
respectively during shear in various Ko consolidated undrained shear strength tests as
shown in Error! Not a valid bookmark self-reference.. According to Ladd and DeGroot
(2003), each type of shear test will achieve a different combination of the two variables.
For example, in triaxial compression, b = 0 and δ = 0o and in triaxial extension, b = 1 and
δ = 90o.

Note: TC is triaxial compression, TE is triaxial extension, DSS is direct


simple shear, PSC is plane strain compression, PSE is plane strain
extension, TTA is true triaxial apparatus, TSHC is torsional shear
hollow cylinder, DSC is directional shear cell, Pi is inner pressures,
and Po is outer pressures
Figure 1: Stress Systems Achievable by Various Shear
Devices for CKoU Testing (modified from Germaine 1982),
from Ladd (1991) as published in Ladd and DeGroot (2003)

2.3 Practical Relevance of Laboratory Strength Testing


Kulhawy and Mayne (1990) discuss the practical relevance of strength tests to
various field conditions, which is summarized in Figure 2. Various field situations are
best represented by the use of different boundary conditions. In a 2-D slope stability
problem for example (Figure 2a), the shear strength at the base of a slip circle is best

5
captured in direct simple shear, that near the toe is best represented by plane strain
extension, and that near the crest is best represented by plane strain compression.
Since triaxial testing provides more conservative strengths than plane strain testing
(Table 1) and since triaxial testing is more practical than plane strain testing, Figure 2a,
stipulates the use of either type of test at the slip surface extremes. Also shown in
Figure 2 are other geotechnical structures with the appropriate lab tests relevant to the
problems.

A comparison of effective friction angles among various shearing modes is shown


in Table 1. According to Kulhawy and Mayne (1990), most soils will be under anisotropic
stress conditions in situ. Consolidating samples anisotropically, as opposed to
isotropically, result in slightly lower effective friction angles, as shown in Figure 3. Also,
anisotropic consolidation can result in lower undrained shear strength than isotropic
consolidation, as shown in Figure 4.

Figure 2: Relevance of Strength Tests to Various Field Practices,


from Kulhawy and Mayne (1990)

6
Table 1: Relative Effective Friction Angles of Normally Consolidated Tests, from
Kulhawy and Mayne (1990)

Figure 3: Comparison of Effective Friction Angles Between Isotropic and Anisotropic


Undrained Compression Tests on Normally Consolidated Soils, data from Mayne
(1985) and Nakase and Kamei (1988), as published in Kulhawy and Mayne (1990)

7
Figure 4: Comparison of Undrained Strength Ratios Between Isotropic and Anisotropic
Undrained Compression Tests on Normally Consolidated Soils, data from Mayne (1985),
Nakase and Kamei (1988), and Mayne and Stewart (1988), as published in Kulhawy and
Mayne (1990)

2.4 Strain Compatibility


Ladd and DeGroot (2003) stressed the importance of considering strain
compatibility when selecting average strengths from triaxial compression, direct simple
shear and triaxial extension tests because peak strengths in the various tests occur at
different strains. Typical compression test behavior for soft clays includes peak strength
at a low strain followed by strain softening. Other shearing modes tend to reach peak
strengths at higher strains, as shown in Figure 5. As a consequence, it can be
rationalized that precarious slopes and footings experience a form of progressive failure
prior to collapse. During progressive failure, the average strength mobilized along this
surface will be less than the average of the peak strengths from triaxial compression,
direct simple shear and triaxial extension tests. To account for this, triaxial stress-axial
strain functions are translated into functions of γ (shear strain), and when combined
with direct simple shear stress-shear strain data, a “design” average shear stress, τave,

8
versus γ curve can be derived. Thus, for any shear strain, the design undrained shear
strength can be calculated and the potential failure surface is assumed to have a shear
strength equally averaged among the three shear types (compression, direct simple
shear, extension), as shown in Figure 5.

Figure 5: Example of the Strain Compatibility Technique (after


Koutsoftas and Ladd 1985), from Ladd (1991) as published in
Ladd and DeGroot (2003)

2.5 Stress History And Normalized Soil Engineering Properties (SHANSEP)


Ladd and Foott (1974) explain the Normalized Soil Parameters (NSP) and Stress
History and Normalized Engineering Properties (SHANSEP) concepts, which are
summarized in the following paragraphs. This section pertains only to undrained (short-
term) shear strengths.

Normalized behavior is defined as when similar soils with identical


overconsolidation ratios (OCR) exhibit similar strength and stress-strain behavior when
normalized by their consolidation stresses. After normalized plots are developed for
each OCR, the resulting NSP can be applied to a variety of stress conditions. The most
regularly used type of parameter in strength analysis is undrained shear strength
Su
normalized by vertical consolidation stress (σ′ ), sometimes also referred to as the
vc

9
𝑐⁄ ratio often mentioned in soil mechanics literature. SHANSEP utilizes the concepts
𝑝
of both stress history and NSP to better determine undrained shear strength
relationships. This method involves the following steps:

1. Develop a stress history profile of a soil of interest using one-dimensional


consolidation testing.

2. Anisotropically consolidate soil samples to 1.5 to 4 times the preconsolidation


pressure. Hold the final consolidation stress long enough for secondary
compression to restore some structure to the soil.

3. Decrease the effective stress acting on the samples to a known


overconsolidation ratio (OCR).

4. Shear the sample with a chosen shearing mode.

5. Repeat steps 2 through 4 for a range of OCR values.

6. Determine the parameters S and m in the SHANSEP equation (Ladd and Foott,
Su
1974) using linear regression of log (σ′ ) and log (OCR):
vc OC

𝑆𝑢 Eq. 2
( ′ ) = 𝑆 (𝑂𝐶𝑅)𝑚
𝜎 𝑣𝑐 𝑂𝐶

where:
𝑆
(σ′𝑢 ) = normalized undrained shear strength for an overconsolidated soil
vc OC

Su = undrained shear strength


σ’vc = vertical consolidation stress
OCR = overconsolidation ratio
m = empirically-determined constant
S = normalized undrained shear strength for the same soil when normally
𝑆
consolidated = (𝜎′𝑢 ) Eq. 3
𝑣𝑐 𝑁𝐶

10
Different relationships will result for each type of test (CKoUC vs. CKoUE vs. CKoUDSS).
The plots representing Equation 2 will be referred to as “SHANSEP curves”.

Ladd and Foott (1974) stated the following advantages of SHANSEP. Use of
SHANSEP can result in cost reduction due to a decrease in number and complexity of
testing. Once a SHANSEP curve has been established for a particular soil, undrained
shear strengths can be obtained from just stress history data. More undrained strength
– stress history data accumulated can be used to further refine the parameters S and m
and fewer or even no strength tests will be required eventually to determine a soil’s
undrained shear strength profile with depth. In addition, when anisotropically
consolidating to 1.5 to 4 times the preconsolidation pressure, sample disturbance is less
of a concern than with traditional testing methods due to slight restructuring of the soil
during virgin consolidation.

Ladd and Foott (1974) also indicate the following limitations of SHANSEP. The
method does not apply well to highly structured soils such as highly cemented soils and
to varved clays because they do not exhibit NSP behavior. The method also does not
apply to irregular soil deposits for which a well-defined stress history cannot be
obtained.

2.6 Definition of Undrained Shear Strength

Ladd and DeGroot (2003) state that the undrained shear strength from CKoUC
and CKoUE tests should be defined as:

𝑆𝑢 = qcos  ′ Eq. 4

where:
𝜎1 −𝜎3
q=
2

‘ = effective friction angle

11
instead of Su = maximum q for an Undrained Stability Analysis because the assumption
of ’ = 0o does not accurately depict the surface of failure orientation when performing
method of slices. Su = qcosφ’ is more conservative than Su = q, and therefore even if this
assumption is incorrect, the calculated Factor of Safety is not overestimated.

Due to the complexity of direct simple shear testing, the state of stress and the
failure orientation are not known. Airey et al. (1985) observed that the failure plane in
drained DSS tests on kaolin was oriented between 5o and 15o to the horizontal.

Ladd and Edgers (1972) also stated that the undrained shear strength from direct
simple shear tests exists somewhere between qf and qfcos. DeGroot et al. (1992)
concluded that the maximum horizontal shear stress from DSS testing gives a decent
estimate of undrained shear strength for undrained stability analyses.

2.7 Previously Measured Undrained Strength Ratios and SHANSEP Research


2.7.1 Triaxial Tests
Ladd et al. (1977) compiled triaxial compression and extension data on normally
consolidated clays prior to 1977 (Table 2). Results from de La Beaumelle (1991)
SHANSEP research on Boston Blue Clay are summarized in Table 3 for triaxial
compression and Table 4 for triaxial extension. Sheahan et al. (1996) researched the
impact of shearing strain rates on undrained strength of resedimented Boston Blue Clay
with OCR values ranging from 1 to 8. A summary of their results is presented in Table 5
for normally consolidated tests and Table 6 for overconsolidated tests. Using test strain
rates of 0.05%, 0.5%, 5%, and 50% per hour, Sheahan et al. (1996) noted that higher
strain rates generally led to greater strength ratios. Bay et al. (2005) determined the “S”
parameter to be 0.32 and “m” parameter to be 0.82 for soft Bonneville clay from Salt
Lake City, Utah after running multiple triaxial compression tests with OCR values ranging
from 1 to 6. Abdulhadi et al. (2012) investigated the impact of different levels of
consolidation stresses on compression behavior of resedimented Boston Blue clay.
Abdulhadi et al. (2012) discovered that the undrained strength ratio decreased with

12
increasing consolidation stresses. A summary of their results are shown in Table 7 and
Table 8.

Table 2: Undrained Strength Ratios from Five Different Normally Consolidated Clays, as
published in Ladd et al. (1977)

Table 3: SHANSEP Parameter Results for Triaxial Compression of Boston Blue Clay, from
de La Beaumelle (1991)

Note: SB is South Boston samples, (T) is Tube samples, (B) is Block samples, NC is Normally
Consolidated, OC is Overconsolidated, SD is Standard Deviation, and USR is Undrained
Strength Ratio

13
Table 4: SHANSEP Parameter Results for Triaxial Extension of Boston Blue Clay,
from de La Beaumelle (1991)

Note: SB is South Boston samples, EB is East Boston samples, (T) is Tube samples, (B) is Block
samples, NC is Normally Consolidated, OC is Overconsolidated, SD is Standard Deviation, and
USR is Undrained Strength Ratio

14
Table 5: Results from Ko-Consolidated Undrained Compression Tests of Resedimented
Normally Consolidated Boston Blue Clay at Various Strain Rates, from Sheahan et al.
(1996)

Table 6 Results from Ko-Consolidated Undrained Compression Tests of Resedimented


Overconsolidated Boston Blue Clay at Various Strain Rates, from Sheahan et al. (1996)

15
Table 7: CKoUC Results of Normally Consolidated Resedimented Boston Blue Clay at
Different Consolidation Stresses, from Abdulhadi (2012)

Table 8: CKoUC Results of Overconsolidated Resedimented Boston Blue Clay at


Different Consolidation Stresses, from Abdulhadi (2012)

16
2.7.2 Direct Simple Shear Tests
Bjerrum and Landva (1966) ran a series of drained and undrained direct simple
shear tests on Norwegian quick clay. This study was one of the first to use a direct
simple shear apparatus from which the concept and mechanisms are the basis of
today’s machines. The apparatus is able to consolidate the sample by applying vertical
dead load, similar to how a traditional consolidometer works. A drawing of the
apparatus is shown in Figure 6. Shearing the sample involves moving the top cap of the
specimen horizontally while keeping the sample height constant. In theory, a constant
sample height results in constant sample volume (which simulates undrained
conditions) assuming that the wire reinforcement prohibits changes in sample cross
section area. Bjerrum and Landva (1966) also theorized that the pore pressure would
be equivalent to the change in vertical pressure required to keep the sample height
constant. Dyvik et al. (1987) verified that this assumption is valid.

Results of constant volume (undrained) simple shear results by Bjerrum and


Landva (1966) are summarized in Table 9. Ladd and Edgers (1972) presented CKoUDSS
SHANSEP curves for five different clays as shown in Figure 7. Kulhawy and Mayne
(1990) found that:

𝑆𝑢 𝑆𝑢 𝑆𝑢
( ) = 0.45 [( ) +( ) ] Eq. 5
σ′vc DSS σ′vc C𝐾 σ′vc C𝐾
𝑜 𝑈𝐶 𝑜 𝑈𝐸

Equation 5 indicates that the undrained strength ratios from CKoUDSS are approximately
the average of CKoUC and CKoUE undrained strength ratios. Therefore, the common
trend is that the CKoUDSS SHANSEP curve plots below the CKoUC SHANSEP curve and
above the CKoUE SHANSEP curve. DeGroot et al. (1992) compiled an extensive amount
of DSS test data on normally consolidated cohesive soils. Figure 8 and Figure 9 show
histograms of normalized undrained strength and shear strain at maximum shear stress
for the accumulated data. Figure 10 shows a plot of shear strain at maximum shear
stress for the accumulated normally consolidated DSS data vs. plasticity index. A plot of

17
normalized undrained strength vs. shear strain at maximum shear stress is shown in
Figure 11. It should be noted that there is considerable scatter in the DeGroot et al.
(1992) data. A comparison of undrained strength ratios between laboratory and field
vane test data from DeGroot et al. (1992) is shown in Figure 12. Although there is
considerable scatter in the field vane data, the laboratory data seem to underestimate
the undrained strength ratio at higher plasticity. DeGroot et al. (1992) also made the
following conclusions on the strain softening behavior of direct simple shear tests on
cohesive soil:

1. Normally consolidated specimens exhibit strain softening behavior when tested


in the Geonor DSS machine.
2. There is little or no correlation between strain softening and plasticity or liquidity
indices.
3. Measured DSS stress-strain behavior is influenced by stress non-uniformity.
Although it is believed that peak shear resistance is not affected much, the tests
exhibit strain softening that is larger than actual soil behavior.
4. If ruptures occur in all DSS tests at peak strengths, then post rupture data are
not representative of in situ soil behavior.

18
Figure 6: Direct Simple Shear Apparatus at Norwegian Geotechnical Institute, from Bjerrum and Landva (1966)

19
Table 9: Results of Constant Volume Direct Simple Shear Tests, from Bjerrum and
Landva (1966)

Figure 7: SHANSEP Curves of Five Different Normally Consolidated Clays, from Ladd
and Edgers (1972)

20
Figure 8: Histogram of Normalized Undrained Strength for 100 Normally
Consolidated CKoUDSS Tests, from DeGroot et al. (1992)

Figure 9: Histogram of Shear Strain at Maximum Shear Stress for 100


Normally Consolidated CKoUDSS Tests, from DeGroot et al. (1992)

21
Figure 10: Shear Strain at Maximum Shear Stress vs. Plasticity Index for Normally
Consolidated CKoUDSS Tests on Cohesive Soil, from DeGroot et al. (1992)

22
Figure 11: Normalized Undrained Strength vs. Shear Strain at Maximum Shear Stress for
Normally Consolidated CKoUDSS Tests on Cohesive Soil, from DeGroot et al. (1992)

23
Figure 12: Comparison of Field Vane and Laboratory Undrained Strength Ratios for Normally Consolidated
Non-Varved Sedimentary Soils, from Ladd (1991), as published in DeGroot et al. (1992)

24
2.7.3 Combined CKoUC, CKoUE and CKoUDSS Data
Ladd and DeGroot (2003) provided SHANSEP curves for a plastic marine clay at
the site of the Atlantic Generating Station (AGS) project off the coast of New Jersey and
Recompression curves for a sensitive marine clay in James Bay, Canada. The curves
along with the respective values of SHANSEP parameters S and m are shown in Figure
13.

Meng and Chu (2011) developed SHANSEP curves for CKoUC, CKoUE, CKoUDSS
and CKoUPS (plane strain) tests for residual soils in Singapore as shown in Figure 14.
According to Meng and Chu (2011), soil classification was silty clay with sand (CL).
SHANSEP parameters were not published, but were estimated based on Figure 14 to be
as follows: S = 0.29, 0.29, 0.22 and 0.22, and m = 0.97, 0.84, 0.91 and 0.59 for CKoUC,
CKoUPS, CKoUDSS and CKoUE, respectively.

Figure 13: OCR vs. Undrained Strength Ratio and Shear Strain at Failure for Ko
Consolidated Undrained Tests: (a) AGS Plastic Marine Clay (PI = 43%, LI = 0.6) via
SHANSEP (Koutsoftas and Ladd 1985); and (b) James Bay Sensitive Marine Clay (PI =
13%, LI = 1.9) via Recompression (B-6 data from Lefebvre et al. 1983) [after Ladd
1991], as published in Ladd and DeGroot (2003)

25
Figure 14: SHANSEP Curves for Various Shear Modes of Singapore Residual Soils,
from Meng and Chu (2011)

26
3 SITE CHARACTERISTICS
Details of the site where the soil samples used in this research originated from are
discussed in the following section. This chapter includes a description of the site
location, geology, soil stratigraphy and soil characteristics.

3.1 Site Location


Soil samples were taken from borings located south of Ke’ehi interchange where
Nimitz Highway and H-1 Interstate Highway merge. The borings were drilled by
Geolabs, Inc. A map of the peninsula site location is shown in Figure 15. Moanalua
Stream and Kalihi Stream occupy the western and eastern sides, respectively of the
peninsula and channel into Ke’ehi Lagoon. Boring 4-A was drilled near the bank of
Moanalua Stream and borings 5-A and 5-C were drilled near the bank of Kalihi Stream.
Approximate boring locations are shown in Figure 16.

H-1 Interstate Highway

H-1 Interstate Highway Kalihi


Stream
Moanalua
Stream Peninsula
Site Nimitz
Highway

Figure 15: Peninsula Site Location on O’ahu, Hawai’i, images taken from Google Earth
(2013)

27
Moanalua Kalihi
Stream Stream

Figure 16: Approximate Boring Locations, image taken from Google Earth (2013)

3.2 Site Geology


Summarized in the following paragraphs is the geologic history of the island of
O’ahu as described by Macdonald and Abbott (1970). Ultimately, the main focus will be
on the area of the sampling location.

O’ahu was primarily formed by the Waianae and Ko’olau shield volcanoes which
today are the two prominent mountain ranges on the island. After volcanic activity
ceased, the island sank and a broad bay formed on the southern shore. Then, coral
barrier reefs developed in the bay, and terrigenous and marine sediments comprising
primarily of silts and sands were deposited.

After approximately 2 million years of sinking, some volcanic activity resumed and
many tuff cones near the Honolulu region were formed. Thereafter, the island
underwent drastic changes in sea level. When the sea level dropped, erosion formed

28
valleys and streams. Several of these streams then merged into larger “master”
streams. When sea levels rose, sediments deposited on the submerged valleys.

When the Salt Lake tuff cone was formed, Moanalua Stream was redirected
towards Ke’ehi Lagoon. Moanalua Stream and Kalihi Stream became the major deposit
channels for Ke’ehi Lagoon and its vicinity, in which thick layers of lagoonal deposit are
present today.

Clemente (1984) reported that the soils in the vicinity of Ke’ehi Interchange were
underconsolidated in 1984. Clemente states that coral sand was dredged from the
ocean and placed as surface fill in the 1930’s and 40’s, which is apparent in borings 5A
and 5C (Appendix A, which classifies the fill as a tannish white silty gravel with sand) but
not in borings 4A and 4B (fill was described as brown clayey silt). The fact that the soil
was underconsolidated is further supported by Figure 17, which shows a plot of the
excess hydrostatic head with depth as reported by Clemente (1984).

In 1955, Cline (1955) provided a soil survey map of the Hawai’i islands. The map
covering the Ke’ehi area is shown in Figure 18, which clearly shows that the peninsula
site did not exist at that time. Therefore, the site was filled sometime after 1955. The
red dotted area shows an estimate of the peninsula site on the 1955 map. The
peninsula was reclaimed between 1955 and 1964 since the peninsula is visible in a map
published by Visher and Mink (1964) as shown in Figure 19.

Visher and Mink (1964) studied groundwater resources in the southern half of
Oahu, and relevant details are summarized as follows. The deposited sediments form a
low-permeable cap (Visher and Mink called it caprock, which is misleading because
these deposits are not rock. It is referred to herein as “cap”) over the more permeable
basaltic aquifers. The cap provides a barrier that prevents movement of fresh water
from the island to the ocean. Along with rainfall, irrigation and spring water, seeped
artesian water from the aquifer contributes to existing water within the cap. The
basaltic aquifer is constantly recharged as rain water continuously seeps down from the
Ko’olau mountain range. If groundwater recharge is greater than the discharge, artesian

29
conditions can develop within the confined aquifer as shown in Figure 20. This is
relevant to the soils used in this study as it will be shown later that the cap is in fact
underconsolidated as a result of the artesian pressures in the basalt aquifer below.
More importantly, the underconsolidation is not a result of the fill since it has been at
least 50 years since the peninsula site was filled in. It is estimated that primary
consolidation would already be complete by now (2014).

Figure 17: Excess Hydrostatic Head with Depth at Ke’ehi Interchange, from Clemente
(1984)

30
Figure 18: 1955 Map Showing the Estimated Peninsula Site Location Near Ke’ehi
Interchange, from Cline (1955)

Figure 19: Map Showing the Peninsula Site Near Ke’ehi Interchange, from Visher and
Mink (1964)

31
Figure 20: Cross-Section View of Groundwater Occurrence in the Southern Half of O’ahu,
Hawai’i. Diagram Courtesy of Honolulu Board of Water Supply, image taken from Seu
(2002): http://legacy.earlham.edu/~seuka/Hydrogeology_page/

3.3 Soil Stratigraphy


Boring logs 4A, 5A and 5C show that the top 10 to 14 feet was described as fill. Fill
for Boring 4A consisted of a brown stiff clayey silt while fill for borings 5A and 5C
consisted of a tannish white medium dense silty gravel. For Boring 4A, lagoonal deposit
consisting of grayish to brownish elastic silt extends from the bottom of the fill down till
the bottom of the boring at a depth of 44 feet. Boring 5A contains a layer of recent
alluvium consisting of black, loose sand extending from 14 ft to a depth of 29 ft, which in
turn is underlain by a lagoonal deposit consisting of elastic silt with little sand. Boring
5A was terminated at a depth of 44 ft. The 10 ft of fill in Boring 5C is underlain by a 5-ft-
thick layer of sandy silt followed by 15 ft of black to dark gray, very loose sand. Below
the sand is approximately 95 ft of lagoonal deposit consisting of dark gray, very soft silt
(described as silty clay CH in the boring log but tests show that the soil is actually MH).
Boring logs provided by Geolabs, Inc. are shown in Appendix A.

3.4 Test for the Presence of Calcium Carbonate


An acid test to detect the presence of calcium carbonate in the lagoonal deposit
was performed closely but not quite in accordance with ASTM D4373-14 (2014). The

32
test comprised of pouring hydrochloric acid on approximately 1 g of dried soil. As seen
in Figure 21a, there was vigorous bubbling and effervescence strongly indicating the
presence of calcium carbonate in the lagoonal deposit. This vigorous effervescence was
absent in the same test on a tropical residual soil (Figure 21b). It can be concluded that
the lagoonal deposit is of coralline (meaning calcareous) origin.

(a) (b)

Figure 21: Test for the Presence of Calcium Carbonate on (a) the Hawai’i Lagoonal
Deposit and (b) a Tropical Residual Soil

3.5 Soil Characteristics


Undisturbed Shelby tube samples from borings 4A, 5A and 5C drilled on the
peninsula were used in this thesis. The following sections describe the index properties
and preconsolidation pressures of the soil samples. Samples with similar index

33
properties from borings 4A, 5A and 5C were selected for testing. Cone penetration test
(CPT) plots are provided in Appendix B.

3.5.1 Index Properties


Grain-size distribution from sieve and hydrometer analyses and results from
Atterberg limit tests are presented in this section. In all index tests, over-drying the soil
was avoided due to the possible presence of Halloysite mixed in with the corraline fines,
which can undergo irreversible changes upon drying.

Grain-Size Distribution
Sieve analyses were run in accordance with ASTM C136-06 (2006) and
hydrometer analyses in accordance with ASTM D422-63(2007)e2. Since the soil mass
was not allowed to dry before sieving, they were instead washed through the sieve
stack. The retained soil on each sieve was then transferred to separate evaporating
dishes and oven dried. Catching soil fines passing through the #200 sieve was difficult
due to the enormous amount of water used to wash the soil through each sieve. Those
fines were instead allowed to drain off and the total dry mass was calculated using the
total moist weight and moisture content determined prior to wet sieving.

The Shelby tube sample from Boring 5A had 88% passing the #200 sieve. All
other samples contained greater than 95% passing the #200 sieve, most of which had
97%-99% passing. Percentage of particles less than 2 μm ranged from 22% to 37%.
Representative grain-size distribution plots for borings 4A, 5A and 5C are shown in
Figure 22.

34
100%

90%

80%

70%
Percent Passing

60%

50% 4A
5A
40%
5C
30%

20%

10%

0%
10 1 0.1 0.01 0.001 0.0001
Grain Size (mm)

Figure 22: Representative Grain Size Distribution Plots for Samples from Borings 4A, 5A
and 5C

Atterberg Limits
Atterberg limit tests were run in accordance with ASTM D4318-10 (2010). Liquid
limits ranged from 74% to 85%, plastic limits ranged from 40% to 45%, plasticity indices
ranged from 30% to 40%, and liquidity indices ranged from 0.35 to 0.76. Table 10
summarizes the depth and index properties for each Shelby tube sample as well as the
type of shear tests ran. Figure 23 shows where the samples plot on a plasticity chart.
All test samples are Elastic silt with high plasticity (MH). Figure 24 shows a plot of
Atterberg limits and moisture content with depth.

35
Table 10: Depths, Index Properties and Shear Tests Ran for Various Shelby Tube Samples
Liquid Plastic Moisture Number of Tests
Depth Tube Plasticity Liquidity
Limit Limit Content
(ft) No. Index (%) Index CKoUC CKoUE CKoUDSS
(%) (%) (%)
B-4A
25-27 74 40 34 66 0.76 1
SH-7
B-4A
30-32 85 45 40 59 0.36 1
SH-9
B-4A
35-37 80 41 39 67 0.67 1
SH-11
B-4A
40-42 78 43 35 63 0.57 1
SH-13
B-5A
40-42 83 49 34 57 0.24 2 1
SH-13
B-5C
50-52 72 42 31 55 0.41 2 3
SH-17

60

50
4A
Plasticity Indiex (%)

40 5A

30 5C

20 A line

U line
10

0
0 20 40 60 80 100
Liquid Limit (%)

Figure 23: Plasticity Chart for All Triaxial and DSS Samples

36
Plastic Limit, Moisture Content and Liquid Limit (%)
0 20 40 60 80 100
0
4A PL and LL
10
4A Moisture Content
20
5A PL and LL
Depth (ft)

30 5A Moisture Content

40 5C PL and LL

5C Moisture Content
50

60

Figure 24: Atterberg Limits and Moisture Contents vs. Depth for Borings 4A, 5A
and 5C

Specific Gravity
One specific gravity test for each boring was run in accordance with ASTM D854-
10 (2010). Specific gravity values along with sample information are presented in Table
11.

Table 11: Specific Gravity Data and Sample Information for Borings 4A, 5A and 5C

Boring Shelby Tube Depths (ft) Specific Gravity (Gs)


4A SH-9 30-32 2.71
5A SH-13 40-42 2.62
5C SH-17 50-52 2.69

3.5.2 Preconsolidation Pressure


Consolidation tests were run in accordance with ASTM D2435/D2435M - 11
(2011). The loading stress increments were approximately 150, 320, 660, 1310, 2620,
5200, 10400 and 20750 psf. The preconsolidation pressures were determined using
Casagrande’s method. Representative plots of percentage strain vs. consolidation
pressure for borings 4A, 5A and 5C are shown in Figure 25. Preconsolidation pressures

37
ranged from 600 psf to 1200 psf. Preconsolidation pressures and calculated effective
overburden pressures are plotted with depth in Figure 26 for Boring 4A and Figure 27
for borings 5A and 5C. Preconsolidation pressures less than overburden pressures
indicate that the soil is underconsolidated. However, as indicated by Visher and Mink
(1964) in Section 3.2 and confirmed by CPT logs in Appendix B (pore pressures in Boring
4A are not hydrostatic), the basalt aquifer below the lagoonal deposit is under artesian
pressure. Therefore, the pore pressures in the lagoonal deposit must be elevated above
hydrostatic values. The elevated pore pressures reduce the calculated effective
overburden pressure.

Consolidation Pressure (psf)


1 10 100 1000 10000 100000
0

5
4A

10

15
5A
Strain (%)

20

25 5C (from DSS
consol)

30

35

40

Figure 25: Sample Strain vs. Consolidation Pressure Plots for Samples from Borings
4A, 5A and 5C

38
'vo and 'vm (psf)
0 500 1000 1500 2000
0

Water Table
5

10

4A
Depth Below Ground Surface (ft)

Preconsolidation
15 Pressure

20
Calculated
Effective
25 Overburden
Pressure
(assuming
hydrostatic
30 conditions)

35

40

45

Figure 26: Preconsolidation Pressure and Effective Overburden Stress vs. Depth for
Boring 4A

39
'vo and 'vm (psf)

0 500 1000 1500 2000 2500 3000


0

Water Table

10

5A
Preconsolidation
Pressure
20
Depth Below Ground Surface (ft)

5C
Preconsolidation
30 Pressure

Calculated
40 Effective
Overburden
Pressure
(assuming
hydrostatic
conditions)
50

60

Figure 27: Preconsolidation Pressure and Effective Overburden Stress vs. Depth for
Borings 5A and 5C

40
4 TRIAXIAL AND DSS TESTING - SPECIMEN PREPARATION
The following sections describe specimen preparation procedures for triaxial and
direct simple shear tests. This involves extrusion from Shelby tubes, sample trimming
and preparation required for setting the sample into the testing machines.

4.1 Soil Extrusion


Soil samples were extruded from 2.8-inch-diameter, 30-inch-long Shelby tubes.
Tubes were pushed 24 inches into the lagoonal deposit and typical recoveries ranged
from 20 to 24 inches. To minimize disturbance during extrusion, tubes were cut to
approximately 7-inch-long sections adequate for both a triaxial and a direct simple shear
specimen. The Shelby tube was cut with a tube cutter. A deburring tool was used to
smooth out the rough inner edges at the tube cut prior to extrusion, as those
protrusions could potentially damage the sides of the sample during extrusion. A
hydraulic extruder was used to extract the soil, as shown in Figure 28. The soil was cut
with a wire saw. The extruded soil length was slightly larger than the required test
specimen height.

Figure 28: Soil Extruder and Extruded Sample

41
4.2 Triaxial Specimen
4.2.1 Trimming and Dimensions
Samples were trimmed down to 2.4-inch-diameter cylinders from 2.8 inches.
The benefit of sample trimming is that the side of the sample, which is usually the most
disturbed part of the sample, is removed.

Samples were placed into a trimming apparatus and a wire saw was used to
shave off thin slivers at a time until the sample resembled a cylinder. The specimen
length was chosen to ensure a height-diameter ratio of at least 2, as suggested by ASTM
D 4767-11 (2011). The trimming apparatus is shown in Figure 29.

Figure 29: Trimming the Sample in the Trimming Apparatus

4.2.2 Filter Strips


Bishop and Henkel (1962) indicated that use of filter strips assist in sample
saturation and accelerate consolidation times. For compression tests, 8 filter strips
were oriented vertically with each end in contact with the top and bottom of the
specimen. For extension tests, 4 filter strips were spiraled around the sample with each

42
end in contact with the top and bottom of the specimen. ¼-inch-wide filter strips were
used for compression and 3/16-inch-wide for extension tests.

The filter strips were placed by lightly wetting the ends and center of the strips
and carefully sticking them to the sides of the sample while ensuring the strips were
evenly spaced. Strips for extension tests were not long enough to spiral continuously
from the bottom to the top of the specimen. As a result, two strips were spliced with a
slight overlap such that the combined strips were in contact with the bottom and top of
the specimen. Figure 30 shows an example of the filter strip placement for compression
and extension tests.

Figure 30: Filter Strip Placement for Compression (left photo) and Extension (right
photo) Tests

4.2.3 Drainage
A porous stone and circular filter paper was placed at each end of the soil
specimen to allow end drainage. Filter paper was trimmed to the same cross-sectional
area as the porous stone. As suggested by ASTM D 4767-11 (2011), porous stones were
bathed in boiling water for approximately 1 hour prior to testing to remove any trapped

43
fines and cooled off to room temperature. The filter paper used was grade 417 supplied
by VWR International.

4.2.4 Rubber Membrane


Bishop and Henkel (1962) explained that rubber membranes are used to
separate the sample from the confining liquid within the cell chamber. The membrane
dimensions were approximately 8.125 inches high, 1.9 inches in diameter and 0.011
inches thick. Membranes were stretched using a membrane expander as shown in
Figure 31.

Figure 31: Membrane Expander and Membrane

4.2.5 Cell Chamber Apparatus


The triaxial cell chamber consists of a base, wall, roof and rods.

The cell base includes a pedestal, three drainage lines, outlets with manual
valves and a cell chamber outlet. One drainage line leads from the pedestal (bottom of
specimen) to an outlet on the cell base. A second drainage line connects the pedestal

44
and top cap (top of specimen). Finally, a third drainage line leads from the top cap to
another cell outlet. Each outlet has manual valves that remain open throughout the test
except for the outlet leading to the top cap, which is kept closed during shear. The cell
chamber outlet leads to an opening on the base to allow drainage of the chamber
water.

The specimen was placed on top of the filter paper overlying a porous stone,
which in turn was on top of the cell base pedestal. The cross-sectional area of the
specimen and pedestal are identical. A top cap of similar area overlain by a porous
stone and another piece of filter paper were placed on top of the specimen. The top
cap has an indentation for the piston to nest in, along with two small holes for drainage
lines. The membrane was installed around the pedestal, specimen, filter papers, porous
stones and top cap. Two tight fitting o-rings snapped onto the pedestal and top cap to
ensure a fully sealed sample.

The cell wall fits into indentations of the base and roof to ensure proper sealing
of the cell chamber. Being incompressible, de-aired water was used to fill the cell
chamber and apply evenly distributed cell pressures around the specimen.

The cell roof contains an opening for the piston to slide in and out freely. For
compression tests, the piston rests in the top cap indentation. For extension tests, the
piston is screwed into the threading of the top cap. A tightening screw locks the piston
in place while the cell chamber is set up within the machines to prevent premature
loading of the sample prior to testing. A nozzle can be locked into an opening in the
roof to allow air pressures to dissipate while the chamber is being filled with water.

The rods secure the roof and the cell base together. They contain a nut at the
bottom and a tightening screw at the top. The rods fit into notches in the cell base and
roof. Once aligned, the cell chamber is sealed by tightening the screws.

Figure 32 shows the complete cell chamber with an installed sample.

45
Figure 32: Complete Cell Chamber with Sample

4.3 Direct Simple Shear Specimen


4.3.1 Trimming and Dimensions
The sample trimming procedure was performed in accordance with the step-by-
step procedure listed in the manufacturer’s (Geonor) manual (1999). The procedural
steps in the manual are provided in Appendix D. A sample cutter and preparation
apparatus provided by Geonor was used. Figure 33 shows the sample trimming
apparatus. Samples were trimmed to approximately ¾-inch (19 mm) high and 2.6-inch-
diameter cylinders (35 cm2 cross-section area). A small razor blade and wire saw were
used to trim excess soil away from the specimen.

46
Figure 33: Sample Trimming Apparatus

4.3.2 Wire-Reinforced Rubber Membrane


The wire-reinforced membrane manufactured by Geonor has a cross-sectional
area of approximately 5.4 in2 (35 cm2). According to Geonor (1999), the wire
reinforcement has a turn frequency of 20 turns per cm of height, and the wire has a
diameter of 0.15 mm, a Young’s modulus of 150 x 106 kPa and a tensile strength of 560 x
103 kPa. Figure 34a shows the wire reinforced membrane. The membrane was installed
around the sample by means of a membrane expander provided by Geonor. The
membrane expander is shown in Figure 34b.

(a) (b)
Figure 34: (a) Wire Reinforced Membrane for DSS Tests and (b) Membrane Expander
and Membrane

47
4.3.3 Sample Assembly
An assembly is set up prior to testing to hold the sample in place. The sample is
positioned between two sets of filter paper and caps with embedded porous stones.
Each cap contains two drainage openings.

The reinforced rubber membrane was installed around the caps and sample with
the middle of the reinforcement section lining up with the sample mid-height. O-rings
were positioned on the top and bottom caps at the edge of the wire reinforcement to
ensure proper sealing. Clamps lock this setup onto a pedestal to prevent movement of
the bottom cap. The sample assembly is shown in Figure 35.

Figure 35: Sample Assembly

48
5 TRIAXIAL AND DSS TESTING – EQUIPMENT, SOFTWARE AND
TEST PROCEDURES

5.1 Triaxial Testing


5.1.1 Equipment
The triaxial test apparatus consists of three components provided by Geocomp:
LoadTracII, FlowTracII and a computer.

LoadTracII
LoadTracII is a load frame with a movable base platen, an S-type load cell (Figure
36) and a Linear Variable Differential Transformer (LVDT) located under the platen
(Figure 37). A base unit displays the machine readings and houses a high speed,
precision micro stepper motor and electrical components connecting the load cell and
LVDT to the computer. A control panel on the base unit allows the user to manually
adjust the platen height. The cell chamber rests on the movable base platen. The load
cell is attached to the top of the reaction frame directly above the center of the platen.
Raising the platen applies compressive loads onto the specimen, while lowering the
platen induces tensile loads. Figure 38 shows the LoadTracII machine.

Figure 36: S-type Load Cell for Triaxial Tests

49
Figure 37: Embedded LVDT Transducer for Vertical Displacements

Figure 38: LoadTracII Machine

50
FlowTracII
FlowTracII provides cell pressure and back pressure and controls pore water
pressure or volume of the specimen. Typically, two FlowTracII machines are needed to
control both the pore pressure within the sample and cell pressure within the chamber.
The machine consists of a water-filled cylindrical tank and a flow pump with a high
speed, precision micro stepper motor that moves the hydraulic piston within the tank.
A control board with a dedicated CPU uses readings from a pressure transducer
mounted at the end of the cylinder to determine what signals should be sent to the
stepper motor.

A panel on the front of the machine allows the user to control various
operations, such as opening and closing drainage valves, filling or draining the tank with
de-aired water and manually adjusting pressures. There is also a supply line coming out
of the front of the machine and an output line coming out of the back. To feed the tank,
the piston expands the tank chamber to induce negative pressures and draw water from
an outside source via the supply line. During the test, the supply line valve is
automatically closed. The output line valve is opened to allow drainage during
consolidation and control sample and cell pressures during shear. Figure 39 shows the
FlowTracII machine.

51
Supply Line

Water Source
Output Line

Figure 39: FlowTracII Machine

5.1.2 Apparatus
The S-type load cell is positioned just above the piston by adjusting the frame
height and platen position. The sample pressure line is inserted into the cell base outlet
leading to the bottom of the sample while the cell pressure line is inserted into the
outlet leading to the cell chamber. To eliminate as much air as possible within the
specimen and drainage lines, all manual valves on the cell base are opened, and a small
pressure is applied to the bottom of the sample. Water and air bubbles are pushed out
of the specimen through the drainage line from the top of the specimen to the outside
of the cell chamber. The complete triaxial apparatus is shown in Figure 40.

For extension tests, an additional attachment connects to the bottom of the S-


type load cell and functions as a piston clamp. To secure the cell chamber to the
LoadTracII platen, ¼ inch bolts are used to fasten the platen to the cell chamber base via
threaded holes in the latter. The addition of the clamp and bolts induces tensile forces
to the sample when the platen is lowered. The extension attachments are shown in
Figure 41.

Detailed triaxial apparatus setup procedures are provided in Appendix C.

52
Figure 40: Triaxial Apparatus

Figure 41: Extension Attachments

5.1.3 Calibration
Loads, displacements and pressures require calibration to convert the units from
the sensors into the respective units of measure (lb, in and psf).

53
True measurements at different machine readings are measured and entered
into the software, which then performs a linear regression between the true
measurements and the machine readings to calculate calibration factors.

A proving ring, pressure gauge and thin metal plates of known height were used
to measure actual loads, pressures and displacements, respectively.

5.1.4 Testing Phases and Software Input


Test-running procedures are further detailed in Appendix C.
Software
Called Triaxial, the Geocomp computer software is used to set parameters, run
tests and record data. The program can be used to control the 4 test phases:
Initialization, Saturation, Ko Consolidation and Shear. The following are descriptions and
inputs for each phase.

Initialization
During the initialization phase, each specimen was subjected to approximately 3
psi of cell pressure and 1 psi of sample pressure, which were held for an hour. This
phase is designed to check for leaks or problems in the system before the major phases
commence. Typically, a leak would be indicated by a simultaneous water volume
decrease in the cell pressure tank and water volume increase in the sample pressure
tank.

Saturation
Saturation of the sample involves monitoring the pore pressure after repeatedly
applying increases of 288 psf to the cell pressure. The program continuously monitors
Skempton’s pore pressure parameter B until it exceeds 0.95 and the cell pressure
exceeds a minimum value of 2880 psf. If B < 0.95 and the cell pressure reaches 12,960
psf, then the user has to decide whether to abort or continue the test.

If the test is aborted, the lines were bled further to remove air bubbles, and the
tests restarted. When restarting the test, the cell and sample pressures were gradually

54
reduced at an even rate to prevent overloading the specimen. Once, the minimum
required B value was lowered to 0.92 because a B-value of 0.95 was not attained after
multiple test restarts.

Consolidation
After saturation, the specimen was Ko consolidated whereby, the cell pressure
was adjusted to keep the sample area constant. After at least 90% consolidation, based
on Taylor’s (1948) square root of time method and after a minimum duration of 24 hrs,
the next consolidation step (or shear phase if in the final increment) commences. Also,
the test will advance to the next step if a maximum duration of 96 hrs is reached.

In the first triaxial compression test, it was planned to use the following stress
increments: 250, 500, 1000 and 2000 psf. However, during the 250 psf increment, the
specimen swelled. As a result, subsequent vertical pressures were increased to 400,
800, 1600 and 3200 psf. Vertical consolidation stresses were also checked to be beyond
the sample preconsolidation pressures before unloading to the intended OCR or before
shearing. To obtain OCR values of 2, 4 and 8, samples were consolidated to 3200 psf
and rebounded to 1600, 800 and 400 psf, respectively.

Shear
The shear phase involved continuous loading (compression) or unloading
(extension) until the specimen axial strain reached 20%. The shear strain rate was set to
0.5% strain per hour, as recommended by ASTM D4767-11 (2011). The maximum load
was set to the capacity of the S-type load cell. For extension tests, the maximum load is
set to the same value as compression tests, but negative instead of positive. To
simulate undrained conditions, sample pore pressures were continuously adjusted
automatically to keep the sample volume constant.

5.1.5 Corrections
Area Correction
Germaine and Ladd (1988) suggest correcting the specimen area during triaxial
shear due to significant radial deformations. The specimens all barreled during triaxial

55
compression testing. Assuming that the barrel is in the shape of a parabola, Germaine
and Ladd (1988) propose the following area correction for undrained triaxial shear:

2
1 √25 − 20𝜀𝑎 − 5𝜀𝑎2 Eq. 6
𝐴𝑐 = 𝐴𝑜 [− + ]
4 4(1 − 𝜀𝑎 )

where

Ac = corrected specimen area


Ao = specimen cross-section area at the start of shear
εa = axial strain

It is worth noting that the Triaxial program uses a different equation to correct the area
during undrained shear. Therefore, the area correction in the program was set to 0 and
area corrections were manually calculated using Equation 6.

In triaxial extension testing, all samples necked. Therefore, Eq. 6 is not


applicable to extension tests. Instead, an area correction assuming the sample deforms
as a right cylinder was applied:

1 − 𝜀𝑣 Eq. 7
𝐴𝑐 = 𝐴𝑜 [ ]
1 − 𝜀𝑎

where

Ac = corrected specimen area


Ao = specimen cross-section area at the start of shear
εv = volumetric strain = 0 for undrained shear
εa = axial strain

56
Filter Strip Correction
According to Germaine and Ladd (1988), filter strips carry some load during
triaxial compression tests. Therefore, load readings are greater than the actual applied
loads on the specimen. ASTM 4767-11 (2011) recommends applying corrections to
deviator stresses during shear:

For axial strains greater than 2%:

𝑃𝑓𝑝 Eq. 8
𝛥(𝜎1 − 𝜎3 ) = 𝐾𝑓𝑝
𝐴𝑐

For axial strains less than 2%, the corrections to deviator stresses are:

𝑃𝑓𝑝
𝛥(𝜎1 − 𝜎3 ) = 50𝜀1 𝐾𝑓𝑝 Eq. 9
𝐴𝑐

where

Δ(σ1 - σ3) = stress correction to be subtracted from deviator stress


ε1 = axial strain in decimal form
Kfp = load per unit length of filter strips
Pfp = perimeter of specimen covered by filter strips
Ac = cross-sectional area of the specimen after consolidation
A Kfp of 0.9 lb/in (1.6 N/cm) was chosen from the range of 1.3–1.9 N/cm suggested by
Germaine and Ladd (1988). The filter strips covered 2 inches of the specimen perimeter
(10 ¼-in-wide strips).

Membrane Correction
Germaine and Ladd (1988) explained that the rubber membrane provides extra
resistance to specimen deformation. The following corrections from ASTM 4767-11

57
(2011) and Germaine and Ladd (1988) are applied to axial and radial stresses during
shear:

4𝐸𝑚 𝑡𝑚 𝜀1
𝛥(𝜎1 − 𝜎3 ) = Eq. 10
𝐷𝑐

If the unstressed membrane diameter is smaller than the specimen diameter:

𝐷𝑐 − 𝐷𝑖𝑚
𝛥𝜎𝑟 = 2𝑡𝑚 𝐸𝑚 ( ) Eq. 11
𝐷𝑐 𝐷𝑖𝑚

where:

Δ(σ1 - σ3) = stress correction to be subtracted from deviator stress


Δσr = stress correction to be added to radial stress
Em = Young’s modulus for the rubber membrane
tm = membrane thickness
ε1 = axial strain in decimal form
Dc = diameter of specimen after consolidation
Dim = unstretched membrane diameter

According to ASTM D4767-11 (2011), the Young’s modulus of the membrane can
be determined by hanging a 0.5-in-wide circumferential strip of membrane, hanging a
known load thereby stretching the membrane and measuring the force per unit strain.
The setup to determine Young’s modulus is shown in Figure 42. The modulus was
calculated by the following equation:

𝐹⁄
𝐴𝑚 Eq. 12
𝐸𝑚 =
𝛥𝐿⁄
𝐿

58
where:

Em = Young’s Modulus of the membrane


F = force applied to stretch the membrane
ΔL = change in length due to applied force
L = unstretched membrane strip length
Am = effective area of the membrane = 2tmws
tm = membrane thickness
ws = width of the membrane strip = 0.5 in

The membrane Young’s modulus was found to be approximately 120 psi.

Figure 42: Setup for Determining Young’s Modulus of the Rubber Membrane

59
Piston and Top Cap Corrections
The piston was greased before triaxial testing. Therefore, piston friction was
assumed to be negligible.

The cell pressure distributed on the top cap does not act over the entire cap area
due to the piston. Triaxial automatically corrects the deviator load by subtracting:

𝛥𝐹𝑑 = 𝜎𝑐 𝐴𝑝 Eq. 13

where:
ΔFd = load correction
σc = cell pressure
Ap = piston Area
The S-type load cell does not account for the weight of the piston and the top
cap. The piston and top cap were weighed in air and then the estimated buoyant forces
were subtracted to estimate their buoyant weights. The buoyant weights were added
to the deviator load. The top cap is fully submerged while the piston was assumed to be
submerged approximately 5 inches in water. Buoyant forces were calculated by
multiplying the water unit weight by estimated volumes of the top cap and submerged
portion of the piston.

5.2 DSS Testing Program


5.2.1 Equipment
Electric motors are used to apply vertical and horizontal loads onto the
specimen. Loads are measured with S-type load cells. Displacements are measured
using LVDT transducers.

The comparator controls and regulates both the vertical and horizontal motors.
The front of the comparator shows a voltage reading each for horizontal load, vertical
load and vertical displacement. There are 3 different modes for motor control: max
speed, manual and comparator. The max speed mode runs the motor at the highest

60
speed when the buttons indicating the direction of loading are pushed. The manual
mode enables the user to control the direction and speed of the motor. The
comparator mode involves automatic regulation of the motors to hold desired readings
constant. The comparator can be used to regulate the vertical and horizontal loads or
the vertical displacement. Upper and lower limits to hold the loads or displacements
are set with buttons under the appropriate voltage reading display. The comparator is
shown in Figure 43.

The direct simple shear box houses the sample assembly and is shown in Figure
44. The base of the shear box contains braces that secure the sample pedestal in place.
The roof is attached to the vertical motor and clamps onto the specimen top cap. The
base of the shear box is attached to the horizontal motor and sits upon a sliding track.

The machine frame holds the motors, load cells, horizontal transducer and shear
box in their respective positions.

Two burettes are used to saturate the top specimen via drainage lines. The
specimen is saturated by flowing de-aired water from one filled burette to another with
a lower water head. The burettes contain manual valves to control the water flow. A
pressure tank filled with carbon dioxide sits near the apparatus to supply the specimen
with carbon dioxide during saturation. The burettes are shown in Figure 45 and the
carbon dioxide tank is shown in Figure 46.

61
Figure 43: Comparator for Direct Simple Shear Tests

Figure 44: Shear Box

62
Figure 45: Burettes for Specimen Saturation

63
Figure 46: Carbon Dioxide Tank for Specimen Saturation

5.2.2 Apparatus
The direct simple shear apparatus, manufactured by Geonor, functions similarly
to the basic apparatus described by Bjerrum and Landva (1966), except loading controls
are motorized.

The vertical motor and load cell are positioned above the specimen. The
horizontal motor and load cell is attached to the sliding base. Shear clamps on the roof
hold the top of the sample assembly in place while the base, which effectively moves
the bottom of the sample, is pushed to apply simple shear on the sample. An LVDT

64
transducer is attached to the sample pedestal to measure vertical displacements of the
roof. Another LVDT transducer is attached to the machine frame to measure horizontal
displacements of the sliding base.

The drainage lines from the carbon dioxide tank and burettes are attached to the
specimen caps.

The transducers and load cells are connected to the comparator from which the
user can control motor movement and speeds.

The DSS apparatus is shown in Figure 47 and Figure 48. Specific details of the
apparatus assembly procedures are described in Appendix D.

Figure 47: Direct Simple Shear Apparatus

65
Figure 48: Specimen in the Direct Simple Shear Apparatus

5.2.3 Calibration
Since the comparator sends voltage readings to the computer, proper calibration
factors must be input into DSSPro to convert voltage to the respective measurements (N
and mm). The load and displacement calibration factors were found in calibration
certificates dated 2000, so the accuracy of the load and displacement calibration factors
was investigated.

The load cell calibration factors were checked by using a recently calibrated
Monogram digital panel indicator manufactured by Omega (Figure 49) and the LVDT
calibration factors by using thin metal plates of known thickness with the LVDT
transducers. The original load and displacement calibration factors were found to be
accurate at the time of the individual sensor calibration for this thesis.

66
Figure 49: Monogram Digital Panel Indicator

5.2.4 Testing Phases and Software Input


Software
The software DSSPro serves as a data recorder at a set sampling rate. The
program translates voltage readings from the comparator to SI units (kilo Pascal for
stress, Newton for load and millimeter for displacement).

Saturation and Consolidation


Specimen saturation is obtained by connecting drainage lines from the top and
bottom of the specimen to burettes filled with de-aired water. The burette connected
to the bottom of the specimen is filled to a height much greater than the burette
connected to the top. As suggested by Mulilis et al. (1975) and Kliewer (1992), filling
the sample with carbon dioxide helps the specimen achieve saturation at lower back
pressures. Kliewer (1992) states that Henry’s Constant for carbon dioxide is much lower
than air, which means that carbon dioxide can dissolve in water under lower pressures.
Carbon dioxide was allowed to seep through the specimen for 30 minutes. When the
water heights in both burettes were equal, the specimen was deemed saturated and
ready for consolidation.

Before consolidating the specimen, calibration parameters must be set in


DSSPro. The readings are then initialized. The vertical motor is switched to comparator
mode and a desired target vertical load is set. The consolidation phase was run in steps

67
with consolidation pressures of 400, 800, 1600 and 3200 psf. The horizontal motor is off
during this phase and the sliding base is locked in place.

Shear
After consolidation, the vertical motor remains in comparator mode but is
switched to regulate vertical position to maintain a constant sample height throughout
the shear phase. The sliding base is unlocked. DSSPro is closed and reset, parameters
are set to match conditions at the end of consolidation, and data recording is started.
The horizontal motor is turned on and switched to manual mode with a shear speed of
5% shear strain per hour, as suggested by ASTM D6528-07 (2007).

5.2.5 Corrections
Horizontal Load Correction
During shear, the horizontal load data exhibited unnatural jumps. In addition,
the shear stress versus shear strain curves did not peak and strain soften as found in
most constant height (undrained) direct simple shear tests as explained in Section 2.7.2.
Instead, the shear stress versus shear strain curves strain-hardened.

When running the horizontal motor without a specimen, loads were picked up
by the load cell as shown in Figure 50. This was unexpected and it is possible that
extraneous dust or particles may be causing friction to develop in the horizontal piston.

To correct for these extraneous loads, two “correction tests” were run without
the presence of a specimen as shown in Figure 50. These horizontal load values were
averaged and subtracted from the horizontal load data of each direct simple shear test
at the same corresponding time. Although the correction test data showed a similar
trend in the two trials, the difference in shear load corrections could have been as high
as 40 N at one point.

68
200
180
160

Horizontal Load (N)


140
120
100 Trial 1
80 Trial 2
60 Average
40
20
0
0 2000 4000 6000 8000 10000
Time (seconds)

Figure 50: Load vs Time Plots for Direct Simple Shear Tests Without a Specimen

Vertical Displacement Problems in DSSPro


Vertical displacements recorded in DSSPro indicated that the specimen height
did not remain constant during shear. This could potentially mean that there is
specimen volume change which would indicate that the direct simple shear tests may
not have been undrained.

Upon further investigation, it was found that the comparator displayed no


change in vertical displacement during shear, while the DSSPro program would
erroneously record non-zero vertical movement. It was later discovered that this
problem occurs when the LVDT readings for horizontal displacement are changing.

For example, during consolidation, when the horizontal displacement does not
change, the vertical readings are accurate. During shear, the vertical displacement
remains constant. Therefore, the horizontal displacement readings during shear were
accurate but the vertical readings were not.

In the very last direct simple shear test, it was noted that both LVDTs provided
inaccurate displacements. To circumvent this, vertical displacement readings were
observed from the comparator and manually recorded during consolidation. The
comparator does not display horizontal displacements. A new set of calibration factors

69
were determined for the horizontal LVDT prior to testing to manually relate the actual
horizontal displacements to those from DSSPro during shear.

Shearing Rate Problems


The shear speed can be set on the comparator and was estimated by observing
the displacements over time. A perfect shear speed of 5% shear strain per hour was not
always easy to achieve. Actual shear speeds used varied from 4.5% to 6.5% shear strain
per hour. As mentioned in Section 2.7.1, higher strain rates lead to higher strength, but
since the strain rate was only slightly off, the measured undrained strengths should be
accurate enough for all intent and purposes.

70
6 RESULTS
The strength anisotropy due to initial stress system of the lagoonal deposit from
Ke’ehi Interchange is studied by deriving the SHANSEP curves from CKoUC, CKoUE and
CKoUDSS tests. Details of the undrained strength and specimen deformation behavior,
as well as SHANSEP parameters are presented in this chapter.

6.1 Triaxial (CKoUC and CKoUE)


Four CKoUC and four CKoUE tests were conducted at approximately the following
OCRs: 1, 2, 4 and 8. Plots of q/σ’vc vs axial strain are shown in Figure 51 for CKoUC tests
and Figure 52 for CKoUE tests where q = half the failure deviator stress and σ’vc =
effective vertical stress at the end of consolidation. The normalized peak stresses show
an increasing trend with OCR. All triaxial tests show strain softening behavior after the
peak stress.

2.5

OCR 1
q/'vc

1.5 OCR 2
OCR 4

1 OCR 8
Peak

0.5

0
0 5 10 15 20
Axial Strain (%)

Figure 51: q/σ’vc vs Axial Strain Plots for CKoUC Tests

71
-2.5

-2

-1.5
OCR 1
OCR 2
-1 OCR 4
q/'vc

OCR 8
Peak
-0.5

0 -5 -10 -15 -20


0

0.5
Axial Strain (%)

Figure 52: q/σ’vc vs Axial Strain Plots for CKoUE Tests

Peak strength data, maximum obliquity data and sample parameters for all
CKoUC and CKoUE tests are compiled and presented in Table 12. Characteristics such as
deformation behavior, effective stress paths, undrained strength ratios and pore
pressure parameters are also described in the following sections.

72
Table 12: Peak Strength Data, Maximum Obliquity Data and Sample Parameters for All
Triaxial Tests

Peak Max Obliquity


Depths '
Test Sample OCR Kc f (%) q/’vc p’/’vc Af f (%) q/’vc p’/’vc Af
(ft) (deg)
B-4A
25 -27 1.00 0.791 7.70 0.466 0.714 0.754 7.70 0.466 0.714 0.754 40.7
SH-7
B-4A
30 -32 1.98 1.02 3.15 0.728 1.222 0.360 8.25 0.708 1.125 0.423 26
SH-9
CKoUC
B-4A
35 -37 3.96 1.01 7.60 1.422 2.160 0.101 7.60 1.422 2.160 0.101 26
SH-11
B-4A
40 -42 7.71 0.989 4.25 2.401 3.208 0.046 1.30 1.850 2.399 0.129 26
SH-13
B-5A
40 -42 1.00 0.585 -11.30 0.308 0.475 0.814 -8.85 0.297 0.448 0.848 40.4
SH-13
B-5C
50 -52 1.98 0.742 -8.95 0.597 0.454 0.791 -3.50 0.473 0.473 0.836 26
SH-17
CKoUE
B-5C
50 -52 3.88 1.29 -11.00 1.539 0.969 0.563 -1.20 0.925 0.935 0.639 26
SH-17
B-5A
40 -42 7.48 1.53 -6.65 2.005 1.844 0.331 -6.45 1.912 1.915 0.300 26
SH-13

Note: OCR is overconsolidation ratio, Kc is Ko at the end of consolidation, f is axial strain


at failure, q = 0.5(’’p’ = 0.5(’1 + ’3), vc is vertical consolidation stress, Af is
Skempton’s pore pressure parameter = (u - ’)/d, u is pore pressure, d is deviator
stress = ’’ ’ is effective friction angle, ’1 = major principal stress, and ’3 =
minor principal stress.

6.1.1 Deformed Sample Shapes During Shear


For CKoUC tests, the specimens barreled during shear as explained in Section
5.1.5. For CKoUE tests, the specimens necked during shear, which frequently occurred in
the top half of the specimen. Deformation behavior for both CKoUC and CKoUE tests are
shown in Figure 53.

73
Figure 53: Deformation Behavior for CKoUC (left photo) and CKoUE Tests (right photo)

6.1.2 Maximum Obliquity


Failure in triaxial tests can be defined to occur at:

1. Maximum deviator stress (2q)


2. Maximum obliquity (’’or
3. A prescribed value of strain

In this section, failure of maximum obliquity is examined. Plots of normalized ’1/’3 vs


axial strain are shown in Figure 54 for CKoUC tests.

In CKoUC tests, maximum obliquity and peak q occur at the same axial strain for
the OCR = 1 and OCR = 4 tests. For the OCR = 2 specimen, the axial strain at maximum
obliquity is higher than the strain at peak q. For OCR = 8, the axial strain at maximum
obliquity is lower than the strain at peak q.

74
Maximum obliquity for all CKoUE tests occurs at smaller axial strains when
compared to peak q. Since values of ’1/’3 are rather large, especially for some of the
overconsolidated samples, normalized ’1/’3 vs axial strain plots are not presented.
For this thesis, failure was assumed to occur at the more conventionally assumed
maximum q because it resulted in more reasonable and less scattered undrained
strength ratios that are used to develop the SHANSEP curves as discussed in Section 6.3.

0.02

0.018

0.016

0.014

0.012
OCR 1
0.01
('1/'3)/'vc

OCR 2
0.008 OCR 4
0.006 OCR 8
Max Obliquity
0.004

0.002

0
0 5 10 15 20
Axial Strain (%)

Figure 54: Normalized ’1/’3 vs Axial Strain Plots for CKoUC Tests

6.1.3 Effective Stress Paths and Determining φ’


Normalized effective stress paths for all triaxial tests are shown in Figure 55.
Note points at axial strains larger than those at maximum q were truncated for clarity
and to avoid making the figure overly busy. The Kf line is equivalent to the Mohr-
Coulomb failure envelope in p’-q space. It has the following equation:

75
𝑞 = 𝑝′ 𝑡𝑎𝑛𝛼 + 𝑚 Eq. 14

where:

(p’, q) are coordinates of the top of the effective stress failure Mohr-circle

q = 0.5 (’1 - ’3)


p’ = 0.5 (’1 + ’3)
’1 = major principle effective stress
’3 = minor principle effective stress
 = angle of inclination of Kf line with respect to the horizontal = tan-1(sin ’)
’ = effective friction angle of the soil
m = c’cos’
c’ = effective cohesion
When OCR = 1, c’ = m = 0. Therefore, Eq. 14 simplifies to:

𝑞 = 𝑝′ 𝑡𝑎𝑛𝛼 Eq. 15

Hence, it follows from the definition of and Eq. 15 that for normally consolidated soils:

𝑞 Eq. 16
′ = sin−1 ( )
𝑝′

The Kf line for the normally consolidated soil is constructed from the origin to the
failure point (maximum q). The resulting normally consolidated φ’ was 40.7o for CKoUC
tests and 40.4o for CKoUE tests, which is large for a cohesive soil. A possible explanation
for this is the presence of calcium carbonate, as explained earlier in Section 3.4.
Normally, determining the Kf line for overconsolidated soils requires at least 2 tests.
However, due to a limited number of samples, the Kf lines for the overconsolidated soils
are all drawn from the failure points of the respective stress paths (maximum q) at one
constant angle α. The largest angle α that gave the most reasonable set of failure
envelopes was 23.7o (ϕ’ = 26o). Resulting failure envelopes are shown as gray lines in

76
Figure 55. A “reasonable” set of envelopes require that the envelopes do not cross the
q/’vc = 0 axis and the resulting m/’vc values increase with increasing OCR. Also, m/’vc
= q/’vc when p’/’vc = 0.

2.5

1.5

0.5
q/'vc

0
0 0.5 1 1.5 2 2.5 3 3.5
-0.5 OCR 1 Comp
OCR 2 Comp
-1 OCR 4 Comp
OCR 8 Comp
OCR 1 Ext
-1.5
OCR 2 Ext
OCR 4 Ext
-2
OCR 8 Ext

-2.5
p'/'vc

Figure 55: Effective Stress Paths and Failure Envelopes for All Triaxial Tests

Note: Dark gray lines are normally consolidated Kf lines and light gray lines are
overconsolidated Kf lines.

77
6.1.4 Skempton’s Pore Pressure Parameter
Skempton’s pore pressure parameter A can be calculated as:

𝛥𝑢 − 𝛥𝜎′3 Eq. 17
𝐴=
𝛥𝜎′1 − 𝛥𝜎′3

where:

Δu = change in pore pressure


Δσ’1 = change in major principle stress (vertical stress for compression, horizontal stress
for extension)
Δσ’3 = change in minor principle stress (horizontal stress for compression, vertical stress
for extension)

Plots of Skempton’s pore pressure parameter A vs axial strain are shown in Figure
56 for CKoUC tests and Figure 57 for CKoUE tests. At failure (maximum q), Skempton’s A-
parameter is denoted as Af and is plotted vs OCR. The plots are compared with those
from Simons (1960) and de La Beaumelle (1991), as shown in Figure 58. For both CKoUC
and CKoUE tests, Af decreases with increasing OCR. At the same OCR, Af for CKoUE tests
are higher than those for CKoUC tests.

The Af vs OCR plot for CKoUC on Hawai’i lagoonal deposit is very similar to that of
the Boston Blue Clay from de La Beaumelle. Compared to Af vs OCR data for other soils
(Figure 58), the Af values for the Hawai’i lagoonal deposit in CKoUC tests match well at
low OCRs, but are larger at higher values of OCR. For extension, Af values for the
Hawai’i lagoonal deposit are consistently lower than those for Boston Blue Clay.

78
1.0
0.9
0.8
OCR = 1
Pore Pressure Parameter A

0.7
OCR = 2
0.6
OCR = 4
0.5
OCR = 8
0.4
Af Peak
0.3
Af Max
0.2 Obliquity
0.1
0.0
0 5 10 15 20
Axial Strain (%)

Figure 56: Pore Pressure Parameter A vs Axial Strain Plots for CKoUC Tests

1.6

1.4

1.2
OCR = 1
Pore Pressure Paramater A

1 OCR = 2
0.8 OCR = 4

0.6 OCR = 8

0.4 Af Peak

0.2 Af Max
Obliquity
0
0 -5 -10 -15 -20
Axial Strain (%)

Figure 57: Pore Pressure Parameter A vs Axial Strain Plots for CKoUE Tests

79
1.4

1.2
Hawai'i Silt CKoUC
Pore Pressure Parameter at Failure (Af)
1
Hawai'i Silt CKoUE
0.8
Oslo Clay CKoUC
0.6

Oslo and London


0.4 Clay CIUC

Weald Clay CIUC


0.2

0 Boston Blue Clay


CKoUC

-0.2 Boston Blue Clay


CKoUE

-0.4
1 10
Overconsolidation Ratio

Figure 58: Pore Pressure Parameter at Failure (Af) vs OCR for the Hawai’i Lagoonal
Deposit and Other Clays (Superimposed with Oslo, Oslo and London and Weald clays
after Simons, 1960 and Boston Blue clay after de La Beaumelle, 1991)

6.2 CKoUDSS
Plots of horizontal shear stress normalized with the vertical stress at the end of
consolidation vs shear strain for all direct simple shear tests are shown in Figure 59. It is
assumed that failure occurs when the horizontal shear stress is a maximum (τh). The
undrained shear strength Su = τh max. The normalized maximum horizontal shear stress
or undrained strength ratio increases with increasing OCR.

Table 13 presents the compiled results from the four CKoUDSS tests.

80
OCR 1
1.4

1.2
OCR 2
1
OCR 4
0.8
/'vc

0.6
OCR 8
0.4

0.2 Max
Shear
Stress
0
0 5 10 15 20 25
Shear Strain (%)

Figure 59: Normalized Shear Stress vs Shear Strain Plots for All Direct Simple Shear Tests

Table 13: Maximum Shear Stress Data and Sample Parameters

At Maximum h
Sample Depths (ft) OCR (%) h/'vc v/'vc u/'vc h/v
B-5C SH-17 50-52 1 8.15 0.346 0.727 0.329 0.476
B-5C SH-17 50-52 2 18.59 0.769 0.67 0.318 1.147
B-5C SH-17 50-52 4 10.38 1.283 1.874 -0.747 0.685
B-5A SH-13 40-42 8 5.11 1.358 2.965 -1.149 -5.29
Note: his shear stress, is shear strain, v is the applied vertical stress, 'vc is vertical
stress at the end of consolidation.

6.2.1 Pore Pressure


Pore pressures normalized with the vertical stress at the end of consolidation are
plotted vs shear strain for all direct simple shear tests in Figure 60. Note that the pore
pressures in CKoUDSS tests = change in vertical stress from the value at the end of
consolidation. The normally consolidated test shows an increase in pore pressure and

81
therefore, a decrease in vertical stress as the specimen was subjected to horizontal
shear. In the sample with OCR = 2, the pore pressures decrease to negative values at
first and then increase to positive pore pressures at failure. The other overconsolidated
samples show decreases in pore pressures to negative values at failure or increases in
vertical stresses. There are sharp undulations in the normalized pore pressure vs shear
strain curves, but most importantly, the overall trends with OCR appear reasonable.
1
OCR 1
0.5
OCR 2
0
0 5 10 15 20 25 OCR 4
u/'vc

-0.5

OCR 8
-1

Max
-1.5
Shear
Stress
-2
Shear Strain (%)

Note: u/’vc = -v/’vc


Figure 60: Normalized Pore Pressure vs Shear Strain for All Direct Simple Shear Tests

6.3 SHANSEP Curves


A SHANSEP curve for each type of test is constructed using linear regression of
log undrained strength ratio vs log OCR. Each plot contained at least one test on a
normally consolidated soil and at least two on overconsolidated specimens. SHANSEP
curves for CKoUC, CKoUE and CKoUDSS are shown in Figure 61. The OCR = 8 CKoUDSS
and OCR = 4 CKoUE tests were omitted, as their undrained strength ratios appeared to
be outliers. The SHANSEP curve is highest for CKoUC, lowest for CKoUE and in the middle

82
for CKoUDSS. As mentioned in Section 2.6, undrained strength ratios should be defined
as Su = q cos ϕ’ for triaxial tests where q = twice the deviator stress at failure and ϕ’ ≃
40 and 26 for the normally and over-consolidated samples, respectively. Interpreting
undrained strength ratios with this definition results in the triaxial SHANSEP curves
shifting down slightly, as shown in Figure 62.

10

Compression
y = 0.447x0.820
R² = 0.995

DSS
Su/'vc

1 y = 0.371x0.860
R² = 0.959

Extension
y = 0.311x0.929
R² = 1.000

0.1
1 Overconsolidation Ratio 10

Note: Regression equations are in the form of (Su/’vc) = S(OCR)m, i.e. y = Su/’vc and x =
OCR. The OCR = 4 CKoUE and OCR = 8 CKoUDSS data points are not included in the linear
regressions (outliers).
Figure 61: SHANSEP Curves for CKoUC, CKoUE and CKoUDSS Tests with Su = Maximum q
for Triaxial and Su = Maximum τh for CKoUDSS

83
10

Compression
y = 0.357x0.895
R² = 0.999

DSS
Su/'vc

1 y = 0.371x0.860
R² = 0.959

Extension
y = 0.248x1.000
R² = 0.995

0.1
1 Overconsolidation Ratio 10

Note: Regression equations are in the form of (Su/’vc) = S(OCR)m, i.e. y = Su/’vc and x =
OCR. The OCR = 4 CKoUE and OCR = 8 CKoUDSS data points are not included in the linear
regressions (outliers).

Figure 62: SHANSEP Curves for CKoUC, CKoUE and CKoUDSS Tests with Su = q cos ’ for
Triaxial and Su = τh for CKoUDSS

The S and m values for the various tests along with SHANSEP parameters for
other soils are summarized in Table 14. The S values for the lagoonal deposit are the
largest compared to published values for temperate soils and one tropical residual soil
as shown in Section 2.7.3. The presence of calcareous fines in the lagoonal deposit, as
indicated earlier in Section 3.4, provides a possible explanation for the large SHANSEP
parameter S.

84
Table 14: Comparison of S and m values for the Hawai’i lagoonal deposit with those of
other soils

Failure SHANSEP Parameters


Soil Definition for CKoUC CKoUDSS CKoUE
Triaxial S m S m S m
AGS Marine Su = qmax 0.33 0.78 0.26 0.78 0.20 0.86

James Bay Marine Su = qmax 0.45 0.87 0.29 0.70 0.24 0.82

Boston Blue Clay Su = qmax 0.28 0.68 0.20 0.80 0.15 0.86

Singapore Residual Not Defined 0.29 0.97 0.22 0.91 0.22 0.59
Hawai’i Lagoonal 0.45 0.37 0.31
Su = qmax 0.82 0.86 0.93
Deposit
Hawai’i Lagoonal
Su = qmaxcos’ 0.36 0.89 0.37 0.86 0.25 1.00
Deposit

6.4 SHANSEP Su vs Su from Unconsolidated Undrained Triaxial Tests


A comparison was made between the Su calculated from the CKoUDSS SHANSEP
equation and the Su obtained from 11 unconsolidated undrained (UU) triaxial tests
performed on the same lagoonal deposit. UU tests were run in accordance with ASTM
2850-03a (2007). A summary of the comparison is shown in Table 15. It can be seen
that some samples had preconsolidation pressures (’vm) less than the overburden
pressure calculated assuming hydrostatic ground water conditions implying they appear
“underconsolidated” due to the artesian conditions in the basalt aquifer. It is theorized
that the basaltic artesian water pressures have increased the pore pressures in the
lagoonal deposit above hydrostatic values. This has been going on for at least 50 years
(Visher and Mink’s publication is dated 1964) if not for centuries. The samples that are
not underconsolidated are probably near sand layers or lenses that drain the artesian
water.

In these “underconsolidated” samples, the effective vertical stress was taken to


be equal to the preconsolidation pressure and an OCR of 1 was assumed. Therefore,

85
’vo = ’vm and Su = 0.37*’vm for these samples. A plot of Su from SHANSEP vs Su from
UU is shown in Figure 63.

Overall the CKoUDSS SHANSEP equation yielded Su values that are within on
average 10% of those from UU tests, which is remarkably close. It is also worth noting
that SHANSEP gave slightly more conservative Su values compared to UU tests, which is
expected since the strengths in the vertical direction are highest compared to those in
other directions.

Table 15: Comparison Between Su Calculated Using the CKoUDSS SHANSEP Equation and
Su Obtained from Unconsolidated Undrained Triaxial Tests

Su
Shelby Depths 'vm1 Su UU2 SHANSEP - %
Boring OCR SHANSEP
Tube (ft) (psf) (psf) UU (psf) Difference
(psf)
SH-3 10-12 835 1.52 291.4 307.4 -16.0 -5.5%

SH-5 15-17 856 1.20 308.7 377.8 -69.2 -22.4%

SH-7 20-22 731 1.00 270.5 332.5 -62.0 -22.9%


4A
SH-9 25-27 1274 1.00 471.4 417.9 53.5 11.3%

SH-11 30-32 647 1.00 239.4 285.1 -45.7 -19.1%

SH-13 35-37 1441 1.00 533.2 516.7 16.5 3.1%

SH-2 10-12 856 1.56 297.7 320.4 -22.7 -7.6%

SH-5 15-17 940 1.00 347.8 323.9 23.9 6.9%

SH-7 20-22 731 1.00 270.5 422.3 -151.8 -56.1%


4B
SH-9 25-27 898 1.00 332.3 370.9 -38.7 -11.6%

SH-11 30-32 1044 1.00 386.3 466.0 -79.7 -20.6%

SH-13 35-37 982 1.00 363.3 434.6 -71.3 -19.6%


Average3 -28.3 -9.8%

Notes:
1. 'vm = preconsolidation pressure from oedometer samples (Appendix E)
2. Su UU = undrained shear strength from unconsolidated undrained triaxial tests
(Appendix F)
3. Boring 4B SH-7 was an outlier and hence, was not included in average
calculations

86
600

500 y = 1.0324x - 40.544


R² = 0.7486
Su - SHANSEP (psf)

400

300

Outlier Point
200

100

0
0 100 200 300 400 500 600
Su - UU Tests (psf)

Note: The outlier point was not included in the linear regression.

Figure 63: Su Calculated from SHANSEP vs Su Obtained from UU Tests

87
7 SUMMARY AND CONCLUSIONS

7.1 Summary
Shelby tubes of lagoonal deposit were sampled from the banks of Moanalua
Stream and Kalihi Stream, which border a peninsula near Ke’ehi Lagoon on the island of
Oahu, Hawai’i. The lagoonal deposit was determined to be an elastic silt (MH) from
index tests. Based on a hydrochloric acid test, it was determined that the lagoonal
deposit contains calcareous fines chemically comprising of calcium carbonate. After
determining preconsolidation pressures, it was observed that the lagoonal deposit was
underconsolidated.

Four CKoUC tests, four CKoUE tests and four CKoDSS tests were performed and
SHANSEP curves were derived to study the effects of strength anisotropy due to initial
shear stress. Undrained strength ratios were plotted against OCR (overconsolidation
ratio) and SHANSEP parameters for each type of shear test were determined using
linear regression. Undrained strengths were calculated using the CKoUDSS SHANSEP
equation and compared to undrained strengths from UU tests on the same lagoonal
deposit.

7.2 Conclusions
The following conclusions can be made:
1. If Su = q, the SHANSEP parameters are S = 0.447 and m = 0.820 for CKoUC tests S
= 0.321 and m = 0.929 for CKoUE tests.

2. If Su = qcosϕ, the SHANSEP parameters are S = 0.357 and m = 0.895 for CKoUC
tests and S = 0.248 and m = 1.00 for CKoUE tests.

3. For Su = h, the SHANSEP parameters from CKoUDSS tests are S = 0.371 and m =
0.861.

4. If Su = q for triaxial tests, the SHANSEP curves plotted with compression


highest, extension lowest, and direct simple shear in between as expected.

88
However, all SHANSEP S parameters are much greater in comparison to those
for normally consolidated soils presented in Section 2.7.

5. If Su = qcosϕ for triaxial tests, both the CKoUC and CKoUE SHANSEP curves shift
down such that the CKoUDSS SHANSEP curve practically coincides with the
CKoUC curve.

6. The CKoUC, CKoUDSS and CKoUE SHANSEP S parameters fall on the high end of
those presented in Section 2.7, which can be attributable to the presence of
calcium carbonate in the lagoonal deposit as discussed in Section 3.4.

7. CKoUC samples barreled near mid-height while CKoUE samples necked near the
top during shear.

8. Failure envelopes for the normally consolidated samples resulted in an


effective friction angle ϕ’ of about 40o for CKoUC and CKoUE tests. This high
friction angle again can be attributable to the presence of calcareous fines in
the lagoonal deposit. For overconsolidated samples, ϕ’ was estimated to be
26o for both CKoUC and CKoUE tests.

9. Skempton’s pore pressure parameter A at failure (Af) for CKoUE tests are larger
than those for CKoUC tests. Af decreased with increasing OCR for both CKoUC
and CKoUE tests.

10. Pore pressure increased (or vertical stress decreased) for the CKoUDSS test on
the normally consolidated sample during shear. Overconsolidated samples
exhibited decreases in the pore pressure. The OCR = 2 sample failed with
positive pore pressure while the OCR = 4 and OCR = 8 samples failed with
negative pore pressures.

11. The CKoUDSS SHANSEP equation yielded Su values that were within 10% on
average of those from 11 unconsolidated undrained (UU) triaxial tests. The
SHANSEP Su values were more conservative than those from UU tests, which is

89
expected since the strengths in the vertical direction are highest when
compared to those in other directions.

7.3 Recommendations for Future Research


Only one soil in Hawai’i was studied. To get a deeper understanding and
appreciation of strength anisotropy for Hawai’i soils, a wider range of soils should be
tested.

Values of b and are restricted in the CKoUC, CKoUDSS and CKoUE tests as
indicated in Figure 1. For greater flexibility in the values of b and , other types of tests
can be performed such as directional shear cell, true triaxial or plane strain testing.

The specimens tested were not radiographed to assess the quality and uniformity
of the samples. They were assumed to be high quality and uniform but the truth is
unknown. This may be less important with the SHANSEP-type tests but they are
especially important for the oedometer and UU tests.

90
References

Abdulhadi, N. O., Germaine, J. T., and Whittle, A. J. (2012). Stress-dependent Behavior of


Saturated Clay. Canadian Geotechnical Journal, Vol. 49, No. 8, 907-916.
Airey, D. W., Budhu, M., and Wood, D. M. (1985). Some Aspects of the Behavior of Soils
in Simple Shear. Developments in Soil Mechanics and Foundation Engineering, P.
K. Banerjee and R. Butterfield, Eds., Vol. 2, Elsevier, London.
ASTM D2435 / D2435M, 2011, Standard Test Methods for One-Dimensional
Consolidation Properties of Soils Using Incremental Loading, ASTM International,
West Conshohocken, PA. (2011). DOI: 10.1520/D2435_D2435M-11,
www.astm.org.
ASTM Standard C136, 2006, Standard Test Method for Sieve Analysis of Fine and Coarse
Aggregates, ASTM International, West Conshohocken, PA. (2006). DOI:
10.1520/C0136-06, www.astm.org.
ASTM Standard D2850-03a, 2007, Standard Test Method for Unconsolidated-Undrained
Triaxial Compression Test on Cohesive Soils, ASTM International, West
Conshohocken, PA. (2007). DOI: 10.1520/D2850-03AR07.
ASTM Standard D422-63, 2007e2, Standard Test Method for Particle-Size Analysis of
Soils, ASTM International, West Conshohocken, PA. (2007). DOI: 10.1520/D0422-
63R07E02, www.astm.org.
ASTM Standard D4318, 2010e1, Standard Test Methods for Liquid Limit, Plastic Limit,
and Plasticity Index of Soils, ASTM International, West Conshohocken, PA.
(2010). DOI 10.1520/D4318, www.astm.org.
ASTM Standard D4373, 2014, Standard Test Method for Rapid Determination of
Carbonate Content of Soils, ASTM International, West Conshohocken, PA. (2014).
DOI: 10.1520/D4373-14, www.astm.org.
ASTM Standard D4767, 2011, Standard Test Method for Consolidated Undrained Triaxial
Compression Test for Cohesive Soils, ASTM International, West Conshohocken,
PA. (2011). DOI: 10.1520/D4767-11, www.astm.org.
ASTM Standard D6528, 2007, Standard Test Method for Consolidated Undrained Direct
Simple Shear Testing of Cohesive Soils, ASTM International, West Conshohocken,
PA. (2011). DOI: 10.1520/D6528-07, www.astm.org.
ASTM Standard D854, 2014, Standard Test Methods for Specific Gravity of Soil Solids by
Water Pycnometer, ASTM International, West Conshohocken, PA. (2014). DOI:
10.1520/D0854-14, www.astm.org.

91
Bay, J. A., Anderson, L. R., Colocino, T. M., and Budge, A. S. (2005). Evaluation of
SHANSEP Parameters for Soft Bonneville Clays. Utah State University Dept. of
Civil and Environmental Engineering.
Bishop, A. W., and J, H. D. (1962). The Measurement of Soil Properties in the Triaxial
Test. London: Edward Arnold LTDg.
Bjerrum, L., and Landva, A. (1966). Direct Simple Shear Tests on Norwegian Quick Clay.
Geotechnique, Vol. 16, No. 1, 1-20.
Clemente, F. M. (1984). Downdrag, Negative Skin Friction and Bitumen Coatings on
Prestressed Concrete Piles. Ann Arbor, MI: University Microfilms International.
Cline, M. G. (1955). Soil Survey of the Territory of Hawai'i. Soil Survey Series 1939, No.
25. United States Department of Agriculture in cooperation with the Hawai'i
Agricultural Experiment Station.
de La Beaumelle, A. C. (1991). Evaluation of SHANSEP Strength-Deformation Properties
of Undisturbed Boston Blue Clay from Automated Triaxial Testing. thesis
submitted in partial fulfillment of the requirements for the Degree of Master of
Science in civil egineering at the Massachusetts Institute of Technology.
DeGroot, D. J., Ladd, C. C., and Germaine, J. T. (1992). Direct Simple Shear Testing of
Cohesive Soils. Research Report R92-18, Massachusetts Institute of Technology
Center for Scientific Excellence in Offshore Engineering.
Duncan, J. M., and Seed, H. B. (1966b). Strength Variation Along Failure Surfaces in Clay.
JSMFD, ASCE, Vol. 92, No. SM6, 81-104.
Dyvik, R., Berre, T., and Lacasse, S. (1987). Comparison of Truly Undrained and Constant
Volume Direct Simple Shear Tests. Geotechnique, Vol. 37, No. 1, 3-10.
Geocomp. (1982-2005). Triaxial User's Manual. Boxborough, MA.
Geonor. (1999). Instructions for Use of Direct Simple Shear Apparatus.
Germaine, J. T. (1982). Development of the Directional Shear Cell for Measureing Cross-
Anisotropic Clay Properties. thesis presented to the Massachusetts Institute of
Technology, at Cambridge, Massachusetts, in partial fulfillment of the
requirements for the degree of Doctor of Science.
Germaine, J. T., and Ladd, C. C. (1988). Triaxial Testing of Saturated Cohesive Soils. In R.
T. Donaghe, R. C. Chaney, and M. L. Silver, Advanced Triaxial Testing of Soil and
Rock, ASTM STP 977 (pp. 421-459). Philadelphia: American Society for Testing
and Materials.

92
GoogleEarth. (2013, January 29). Ke'ehi Peninsula, 21 deg 19’ 55.73” N 157 deg 53’
29.89” W. Retrieved: October 20, 2014.
Henkel, D. J., and Wade, N. H. (1966). Plane Stain Tests on a Saturated Remolded Clay.
JSMFD, ASCE, Vol. 92, No. SM6, 67-80.
Kliewer, J. E. (1992). Test Procedures for Low-Confining Stress, Multistage Triaxial
Testing of Compacted Cohesive Soils. thesis submitted to Oregon State University
in partial fulfillment of the requirements for the degree of Master of Science.
Koutsoftas, D. C., and C, L. C. (1985). Design Strengths for an Offshore Clay. Journal of
Geotechincal Engineering, ASCE, Vol. 111, No. 3, 337-355.
Kulhawy, F. H., and Mayne, P. W. (1990). Manual on Estimating Soil Properties for
Foundation Design. Final Report (EL-6800) submitted to Electric Power Research
Institute (EPRI), Palo Alto, California.
Ladd, C. C. (1991). Stability Evaluation during Staged Construction (The 22nd Karl
Terzaghi Lecture). Journal of Geotechnical Engineering, Vol. 117, No. 4, 540-615.
Ladd, C. C. (1996). II C: STRESS SYSTEM: Experimental Techniques and Results (Cohesive
Soils), Section 2: Types of Anisotropy. Retrieved from
http://ocw.mit.edu/courses/civil-and-environmental-engineering/1-322-soil-
behavior-spring-2005/lecture-notes/
Ladd, C. C., and DeGroot, D. J. (2003). Recommended Practice for Soft Ground Site
Characterization: Arthur Casagrande Lecture. 12th Panamerican Conference on
Soil Mechanics and Geotechnical Engineering. Cambridge, Massachusetts.
Ladd, C. C., and Edgers, L. (1972). Consolidated-Undrained Direct-Simple Shear Tests on
Saturated Clays. Research Report 72-82, Massachusetts Institute of Technology
Department of Civil Engineering.
Ladd, C. C., and Foott, R. (1974). New Design Procedure for Stability of Soft Clays.
Journal of the Geotechnical Engineering Division, Vol. 100, GT7, 763-786.
Ladd, C. C., Bovee, R. B., Edgers, L., and Rixner, J. J. (1971). Consolidated-Undrained
Plane Strain Shear Tests on Boston Blue Clay. Research Report R71-13, No. 273,
Department of Civil Engineering, Massachusetts Institute of Technology.
Cambridnge.
Ladd, C. C., Foott, R., Ishihara, K., Schlosser, F., and Poulos, H. G. (1977). Stress-
Deformation and Strength Characteristics. Proceedings of the Ninth International
Conference On Soil Mechanics and Foundation Engineering, Vol. 2, (pp. 421-494).
Tokyo, Japan.

93
Larsson, R. (1980). Undrained Shear Strength in Stability Calculation of Embankments
and Foundations on Soft Clays. Canadian Geotechnical Journal, Vol. 17, No. 4,
591-602.
Lefebvre, G., Ladd, C. C., Mesri, G., and Tavenas, F. (1983). Report of the Testing
Committee. Committee of Specialists on Sensitive Clays on the NBR Complex,
SEBJ, Montreal, Annexe I.
Macdonald, G. A., and Abbott, A. T. (1970). Volcanoes in the Sea: The Geology of Hawaii.
Honolulu: University of Hawai'i Press.
Mayne, P. W. (1985). A Review of Undrained Strength in Direct Simple Shear. Soils and
Foundations, Vol. 25, No. 3, 64-72.
Mayne, P. W., and Stewart, E. H. (1988). Pore Pressure Behavior of Ko-Consolidated
Clays. Journal of Geotechnical Engineering, ASCE, Vol. 114, No. 1, 76-92.
Meng, G.-H., and Chu, J. (2011). Shear Strength Properties of a Residual Soil in
Singapore. Soils and Foundations, Vol. 51, No. 4, 565-573.
Mulilis, J. P., Chan, C. K., and Seed, H. B. (1975). The Effects of Method of Sample
Preparation on the Cyclic Stress-Strain Behavior of Sands. Earthquake
Engineering Research Center, University of California, Berkeley, Report No. EERC
75-18.
Nakase, A., and Kamei, T. (1986). Influence of Strain Rate on Undrained Shear
Characteristics of Ko-Consolidated Cohesive Soils. Soils and Foundations, Vol. 26,
No. 1, 85-95.
Seu, K. J. (2002, May 6). The Hydrogeological Resources of O'ahu. Retrieved from
Hydrogeology 2002, earlham college:
http://legacy.earlham.edu/~seuka/Hydrogeology_page/
Sheahan, T. C., Ladd, C. C., and Germaine, J. T. (1996). Rate-dependent Undrained Shear
Behavior of Saturated Clay. Journal of Geotechnical Engineering, Vol. 122, No. 2,
99-108.
Simons, N. E. (1960). The Effect of Overconsolidation on the Shear Strength
Characteristics of an Undisturbed Oslo Clay. Proceedings, Research Conference
on Shear Strength of Cohesive Soils, ASCE, (pp. 747-763).
Taylor, D. W. (1948). Fundamentals of Soil Mechanics. New York: John Wiley and Sons.
Vaid, Y. P., and Campanella, R. G. (1974). Triaxial and Plane Strain Behavior of Natural
Clay. JGED, ASCE, Vol. 100, No. GT3, 207-224.

94
Visher, F. N., and Mink, J. F. (1964). Ground-Water Resources in Southern Oahu, Hawaii.
Washington: United States Government Printing Office.

95
Appendix A: Boring Logs (Geolabs, Inc, 2011)

96
97
98
99
100
101
102
103
104
105
106
107
108
109
Appendix B: CPT Logs (Geolabs, Inc, 2011)

110
111
112
Appendix C: Detailed Testing Procedures for Triaxial Tests
(Triaxial, 2005)

113
114
115
116
117
118
119
120
Appendix D: Detailed Testing Procedures for DSS Tests (Geonor,
1999)

121
Figure 64

122
Figure 65

Figure 66

Figure 67

Figure 68

Figure 69

Figure 70

Figure 71

Figure 72

Figure 73

123
Figure 74

Figure 75
Figure 76

Figure 77

Figure 78

Figure 79

Figure 80
Figure 81

Figure 82

Figure 83

Figure 84

124
125
Figure 85

Figure 86

Figure 87

126
Figure 88

Figure 89

Figure 90

127
128
129
130
Figure 64: Membrane in Expander

Figure 65: Placing Cutter

131
Figure 66: Sample Trimming

Figure 67: Attaching Mounting Ring

132
Figure 68: Placing Lower Filter Holder

Figure 69: Locking Lower Filter Holder in the Mounting Ring

133
Figure 70: Inverting Setup

Figure 71: Removing Disk

134
Figure 72: Mounting Caliper

Figure 73: Lowering Cutter Yolk

135
Figure 74: Mounting Membrane and Expander

Figure 75: Mounting Top Yolk

136
Figure 76: Attaching Upper Filter Holder with Top Yolk

Figure 77: Lowering Upper Filter Holder On Sample

137
Figure 78: Lowering Cutter Yolk

Figure 79: Installing Membrane

138
Figure 80: Membrane in Place

Figure 81: Placing O-rings

139
Figure 82: Weighing Assembly

Figure 83: Attaching Drainage Lines

140
Figure 84: Securing Assembly on Base

Figure 85: Measuring 49 mm Space to Center Sample

141
Figure 86: Lowering Vertical Ram to Barely Touch Sample

Figure 87: Attaching Vertical LVDT

142
Figure 88: Tightening Knobs for Shear Clamps

Figure 89: Shear Clamps in Place

143
Figure 90: Burettes for Saturation

144
Appendix E: Consolidation Stress-Strain Curves

145
Log v (kPa) vm ≈ 40 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 91: Consolidation Stress-Strain Curve for Boring 4A SH-3

146
Log v (kPa) vm ≈ 41 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 92: Consolidation Stress-Strain Curve for Boring 4A SH-5

147
Log v (kPa) vm ≈ 35 kPa
1 10 100 1000 10000
0%

10%

20%
Ԑ%

30%

40%

Figure 93: Consolidation Stress-Strain Curve for Boring 4A SH-7

148
Log v (kPa) vm ≈ 61 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 94: Consolidation Stress-Strain Curve for Boring 4A SH-9

149
Log v (kPa) vm ≈ 31 kPa
1 10 100 1000
0%

10%
Ԑ%

20%

30%

Figure 95: Consolidation Stress-Strain Curve for Boring 4A SH-11

150
Log v (kPa) vm ≈ 69 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 96: Consolidation Stress-Strain Curve for Boring 4A SH-13

151
Log v (kPa) vm ≈ 41 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 97: Consolidation Stress-Strain Curve for Boring 4B SH-2

152
Log v (kPa) vm ≈ 45 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 98: Consolidation Stress-Strain Curve for Boring 4B SH-5

153
Log v (kPa) vm ≈ 35 kPa
1 10 100 1000 10000
0%

10%
Ԑ%

20%

30%

Figure 99: Consolidation Stress-Strain Curve for Boring 4B SH-7

154
Log v (kPa) vm ≈ 43 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 100: Consolidation Stress-Strain Curve for Boring 4B SH-9

155
Log v (kPa) vm ≈ 50 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

Figure 101: Consolidation Stress-Strain Curve for Boring 4B SH-11

156
Log v (kPa) vm ≈ 47 kPa
1 10 100 1000
0%

10%

20%
Ԑ%

30%

40%

50%

Figure 102: Consolidation Stress-Strain Curve for Boring 4B SH-13

157
Appendix F: UU Stress-Strain Curves

158
35

Deviator Stress (kPa) 30

25

20

15

10
Ԑf = 7.07 %
5 'f= 28.4 kPa
Su = 14.7 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14

Axial strain

Figure 103: Stress-Strain Curve for B-4A SH-3 UU Test

40

35

30
Deviator Stress (kPa)

25

20

15

10
Ԑf =5.76 %
5 'f= 36.2 kPa
Su = 18.1 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14

Axial strain

Figure 104: Stress-Strain Curve for B-4A SH-5 UU Test

159
35

30
Deviator Stress (kPa)

25

20

15

10
Ԑf =11.21 %
5 'f= 31.8 kPa
Su = 15.9 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14

Axial strain

Figure 105: Stress-Strain Curve for B-4A SH-7 UU Test

45

40

35
Deviator Stress (kPa)

30

25

20

15

10
Ԑf =5.4
=5.20%%
5 
'f=f=40.0
40.0kPa
kPa
SSu == 20.0 kPa
0 u 20.0 kPa

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14

Axial strain

Figure 106: Stress-Strain Curve for B-4A SH-9 UU Test

160
30

25
Deviator Stress (kPa)

20

15

10

Ԑf =10.99 %
5 Ԑf =13.5 %
' = 27.3 kPa
f=f 27.2 kPa
SS u == 13.6
13.6kPa
kPa
u
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14

Axial strain

Figure 107: Stress-Strain Curve for B-4A SH-11 UU Test

55

50

45

40
Deviator Stress (kPa)

35

30

25

20

15

10 Ԑf =5.30 %
5 'f= 49.5 kPa
Su = 24.7 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Axial strain

Figure 108: Stress-Strain Curve for B-4A SH-13 UU Test

161
35

30

25
Deviator Stress (kPa)

20

15

10
Ԑf = 5.34 %
5 'f = 30.7 kPa
Su = 15.3 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Axial strain

Figure 109: Stress-Strain Curve for B-4B SH-2 UU Test

35

30
Deviator Stress (kPa)

25

20

15

10
Ԑf =4.26 %
5 'f = 31.0 kPa
Su = 15.5 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Axial strain

Figure 110: Stress-Strain Curve for B-4B SH-5 UU Test

162
45

40

35
Deviator Stress (kPa)

30

25

20

15

10 Ԑf =6.85 %
5 'f = 40.4 kPa
Su = 20.2 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Axial strain

Figure 111: Stress-Strain Curve for B-4B SH-7 UU Test

40

35

30
Deviator Stress (kPa)

25

20

15

10
Ԑf =7.94 %
5 'f= 35.5 kPa
Su = 17.8 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Axial strain

Figure 112: Stress-Strain Curve for B-4B SH-9 UU Test

163
50

45

40

35
Deviator Stress (kPa)

30

25

20

15

10 Ԑf =7.11%
5 'f= 46.2 kPa
Su = 23.1 kPa
0
0% 1% 2% 3% 4% 5% 6% 7% 8% 9% 10% 11% 12% 13% 14% 15%
Axial strain

Figure 113: Stress-Strain Curve for B-4B SH-11 UU Test

45

40

35
Deviator Stress (kPa)

30

25

20

15

10 Ԑf =8.71 %
5
'f= 41.6 kPa
Su = 20.8 kPa
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Axial strain

Figure 114: Stress-Strain Curve for B-4B SH-13 UU Test

164

S-ar putea să vă placă și