Sunteți pe pagina 1din 13

Tectonophysics 482 (2010) 29–41

Contents lists available at ScienceDirect

Tectonophysics
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / t e c t o

A plea for more caution in fault–slip analysis


Blanka Sperner a,⁎, Peter Zweigel b
a
Geological Institute, Technische Universität Bergakademie Freiberg, Bernhard-von-Cotta-Str. 2, 09599 Freiberg, Germany
b
Statoil ASA, Postboks 273, 7501 Stjørdal, Norway

a r t i c l e i n f o a b s t r a c t

Article history: The analysis of fault–slip data is a widely used tool in structural geology. This method consists of three major
Received 12 September 2008 steps: (a) data acquisition by measurements in outcrops, (b) data separation into homogeneous subsets and
Received in revised form 9 July 2009 calculation of the principal stress or strain axes with one of several available methods, and (c) data
Accepted 28 July 2009
interpretation in a regional tectonic context, which includes the assignment of ages and the derivation of
Available online 3 August 2009
regional patterns. Each one of these three steps may introduce artefacts and misinterpretations. The major
Keywords:
problem during step (a) is sampling bias due to unfavourable outcrop geometry or due to neglecting faults
Fault-slip analysis with subtle slip–sense indicators. Step (b) contains numerous possibilities to influence the calculation
Stress tensor results, e.g. by arbitrary separation of data into subsets, choice of the analysis method, and uncritical use of
Data documentation software. Age assignment in step (c) can in most cases be made only indirectly, e.g. by the assumption that
measured outcrop-scale faults have the same age as parallel map-scale faults or by the inference that subsets
from different sites with parallel-oriented principal axes are syngenetic. Neither assumption is necessarily
valid. With this paper we want to sharpen researchers' awareness of the problems related to fault–slip
analysis.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction that compression is positive with the principal stress axes being
σ1 ≥ σ2 ≥ σ3 and the stress ratio R = (σ2 − σ3)/(σ1 − σ3).
Faults and fractures are one of the main expressions of deforma- Fault–slip analysis involves, irrespective of the calculation method
tion under brittle conditions, occurring on all scales from microscopic employed, three major steps which are data collection, data analysis,
fractures through map-scale faults to plate boundaries with hundreds and interpretation of the analysed data. Published studies frequently
of kilometres length. The measurement and analysis of mesoscale only document the results of the first two steps and the discussion and
faults, i.e. faults that can be inspected with the naked eye in outcrops, results of the last step. The procedures used during collection and data
and related slip indicators has become very popular since the late analysis are mostly sketched rather vaguely and the data documen-
1970s. Such analyses serve several purposes, e.g.: tation itself does often not easily allow for quality evaluation. Fre-
— to determine the slip direction and slip sense of map-scale faults, quently, the results of data analysis (mainly the orientations of
whose geometry can be constructed by structural mapping techni- principal stress or strain axes) are later uncritically used in regional
ques, sometimes aided by interpretation of seismic sections, whereas studies and quoted as ‘facts’. The same is the case for the inter-
the determination of the movement along these faults is often not pretation step, i.e. particularly the assignment of age intervals to fault
directly possible; sets interpreted as co-genetic, and the derivation of regional patterns
— to determine the regional deformation pattern, i.e. the orien- for such intervals. Especially mean ‘regional’ principal stress or strain
tation of the principal stress or strain axes and the shape of the stress axes for certain age intervals serve then as constraints for tectonic
or strain ellipsoid. models.
There exists an increasing number of methods for interpreting All three major steps contain sources for error, and the degree
fault-slip data and some disagreement prevails concerning the of the interpreter's subjective influence on the result increases from
meaning of the determined principal axes, if they represent strain step to step. We want to make both, users of the method and users of
(kinematics) or stress (dynamics); for a discussion see Twiss and the produced data, aware of some sources of error and of some —
Unruh (1998). We do not intend to get involved with that discussion conscious or unconscious — subjective decisions that are made during
but prefer to use stress nomenclature in the present paper for the fault–slip analysis. Our intention is thus not to argue against the use of
methods we applied are related to dynamics. We use the convention the method, but merely to plead for more caution in its application
and for more criticism in the use of its results.
Below, we will first explain the three major steps of fault–slip
⁎ Corresponding author. Tel.: +49 3731 39 3813; fax: +49 3731 39 3597. analysis in more detail, and we will present for each step some of the
E-mail address: blanka.sperner@geo.tu-freiberg.de (B. Sperner). sources for error and some of the subjective decisions made by the

0040-1951/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2009.07.019
30 B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41

interpreter. Then, we will outline some methodological implications specific rock types). In principle, it is more meaningful to analyse
such as the requirement for more extensive data documentation, and each outcrop separately.
reflect in a meta-methodological manner on the need for critical (d) Misinterpretation of superposed fibres: It is occasionally assumed
appraisal of the way scientific methods are applied in practice. that differently oriented fibres lying upon each other give
relative ages for the related two movements with the fibres on
top reflecting the younger movement. This is an unjustifiable
2. Step 1: Data acquisition in the field assumption, because the crack related to the younger fibre set
not implicitly must develop in front of the older one (i.e. closer
The basic data needed for all methods of kinematic or dynamic to the observer); it also might develop behind it. Which fibres
fault–slip analysis are listed in Table 1. We want to outline some we see lying on top and which thereunder, is only dependent
typical pitfalls during field work: on the block which was accidentally removed during erosion
(Fig. 2).
(a) Unknown slip sense/pseudo-conjugate fault sets: Fault planes (e) Misinterpretation of curved fibres: Doubly-curved fibres with the
with indefinable slip sense are not suitable for the analysis. The two outer sections being parallel seem to reflect three different
casually seen practice of assigning fault planes without known faulting events (e.g. Meschede and Decker, 1993), but they can
slip sense to the best-fitting stress tensor simulates a larger more easily be explained by two events with one producing the
data set, but gives no better quality of the results. The same is middle part and the other one producing the two outer parts
true for seemingly conjugate fault sets without slip sense or (Fig. 3). They even might stem from a single event when move-
even without any lineation. The probability that these faults ment along an irregular plane is temporarily deflected while
belong to different events can never be excluded (Fig. 1). overriding an asperity. For assigning relative ages to these fibres,
(b) Selective data acquisition: Outcrops with a dominant wall it is necessary to distinguish between syntaxial and antitaxial
orientation favour selective data acquisition, e.g. fault planes at fibre growth (Twiss and Gefell, 1990).
obtuse angles to the outcrop wall might be underrepresented,
while others subparallel to the wall tend to be overrepresented in 3. Step 2: data analysis
the measured data set. In this case, concentration on fault planes
in underrepresented orientation (with respect to the outcrop 3.1. Data consistency
wall) might result in data sets with varying orientations being
most suitable for fault–slip analysis (see below). Before starting with the analysis itself, it is appropriate to check
(c) Pooling of data: A prerequisite for meaningful fault–slip analysis the data for consistency to ensure that the measured striations are
is that the displacements in a fault set to be analyzed were within the associated fault planes. Deviations from the fault plane
generated under homogeneous stress (or strain). This entails that might occur from small measurement errors (e.g. uneven fault planes,
it must be ensured that blocks (rock volumes) which contain inaccuracy in compass handling) and are in the order of a few degrees.
faults did not undergo mutual rotations, or these rotations have Appropriate computer programs ‘correct’ these data so that striations
to be quantified (e.g. by fault restoration, see Section 4.2.). In and fault planes are parallel; data with too large deviations are not
cases where outcrops contain major faults, the possibility of block useful for the analysis (e.g. program CHECK; Sperner et al., 1993).
rotation must be considered before either combining all fault–
slip data or grouping them into sets in blocks with assumed 3.2. Separation of data
homogeneous stress (or strain). Likewise, combination of fault–
slip data from different outcrops in direct neighbourhood (a few The two most critical points in fault–slip analysis are the separation
hundred meters distance) to one single data set is only of the faults into homogeneous subsets and the selection of a suitable
reasonable if it can be assured that the outcrops lie within the analysis method for each subset. Liesa and Lisle (2004) tested several
same tectonic unit, that no large fault zone disturbs the stress (or automatic separation methods and came to the conclusion that a fully
strain) field (rotation of the principal axes within short automatic separation procedure is not reasonable. At best, the
distances), that no mutual block rotation occurred between the separation can be done by means of field criteria, e.g. overprinting
outcrops, and that the deformation conditions are the same in relationships, two or more striations on the same fault plane, or different
both places (pre-existing weakness planes might occur only in mineralisation on fault planes. If this is not possible, the only way is a
careful separation by using compatibility criteria and mathematical
analysis. The latter bears a high error potential which can be illustrated
Table 1
with a simple 2D-example in which two data clusters are separated
Parameters and observations to be noted as a prerequisite for sophisticated fault–slip
analysis. mathematically resulting in three artificial, non-realistic clusters (Fig. 4;
see also Lisle and Vandycke, 1996). Similar errors might be introduced
Necessary/obligatory:
during the mathematical separation of fault–slip data with the only
— Fault plane attitude (dip azimuth and dip angle)
— Slip orientation (striae; bearing and plunge) difference that they are analysed in a 4D-space (orientation of the three
— Slip sense (e.g. Doblas, 1998; Hancock, 1985; Petit, 1987) principal stress or strain axes and the ellipsoid's shape ratio). In both
— Confidence level for slip sense determination (e.g.: 1, certain; 2, reliable; 3, cases, plotting of the data provides the opportunity to avoid irrational
unreliable; 4, very unreliable; Hardcastle, 1989) results and to make a rough separation of the data ‘by eye’. For this
Advised:
objective, the most expedient plotting type for fault–slip data is the
— Type of indicator for slip-sense determination
— Relative size of fault plane (in case of large planes) ‘Hoeppener method’ (Hoeppener, 1955): fault planes are plotted as
— Amount of displacement poles in an equal area, lower hemisphere stereonet with an arrow
If applicable: through the pole indicating the slip-direction of the hanging-wall block
— Type and colour of mineralisation on fault plane (→genetic groups)
(computer program F-SH; Sperner et al., 1993). In the following, we
— Age relations between different slip directions on one fault plane
— Age relations between different fault planes
refer to this plotting method as ‘tangent–lineation plot’ (cf. Twiss and
— Bedding attitude Unruh, 1998, who use, however, the slip direction of the footwall block).
— Relation of fault to local structure (e.g. cutting fold, folded fault, and fault Data which are incompatible with each other are easily recognized by
truncated by discordance; if structure varies through outcrop: outcrop sketch having, for example, the same orientation of the fault plane (same pole),
with position of faults indicated)
but different directions of slip (arrows point in different directions;
B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41 31

Fig. 1. Misinterpretation of pseudo-conjugate faults (lower hemisphere, equal area plots). (a) Apparently conjugate faults (field measurements from E-Tibet). In this interpretation
(without known slip direction) the maximum principal stress axis (σ1) would be vertical and the minimum principal stress axis (σ3) NNW–SSE. (b) Two deformation events as
derived from the observed slip directions. Principal stress axes are calculated from the plotted data together with other data from the same outcrop being compatible with the
respective event. For none of the events a vertical maximum principal stress axis (σ1) is revealed and thus interpretation (a) is clearly wrong.

Fig. 5). After this rough separation by graphical means, the calculation of bance of the stress/strain field in the vicinity of faults with large
the reduced stress or strain tensor will already yield a reasonable result displacements or stress/strain refraction due to competence varia-
and the final separation of the data can be done mathematically by tion, might be interpreted as different events which in reality are a
eliminating data with large angles between measured and theoretical single one. Even movements along one and the same fault plane
slip direction. might differ with time and position during one event if rotational
Deformation not necessarily needs to be homogeneous, even on deformation is taken into account (Twiss et al., 1991) or if strain
outcrop scale, but well-separated fault sets always have to be. Thus, partitioning leads to different slip directions being active synchro-
local effects on the orientation of the principal axes, like the distur- nously (Angelier, 1994). Indications for local effects might be, for

Fig. 2. Two generations of fibres on a fault plane. (a) Initial situation. (b) Situation after movement into two directions, strike–slip and down–dip, being active one after another
(sequence not relevant). Both movements result in fibre growth behind steps in the fault plane. (c) Two options of fault plane view after erosion of one of the blocks. The left block
shows down–dip fibres on top, the right block strike–slip ones. This demonstrates that the sequence of fibre layers does not indicate the sequence of slip events.
32 B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41

Fig. 3. Bent fibres on a fault plane caused by two subsequent slip movements, first down–dip, then left-lateral oblique slip. (a) Situation after down–dip slip. (b) Final situation for the
case of antitaxial fibre growth, i.e. fibres grow at the contact with the host rock. (c) Final situation for the case of syntaxial fibre growth, i.e. fibres grow in the middle of the fibre
package.

instance, gradual changes in the orientations of the principal axes or a tesimal strain due to slip on each single fault, but here with an angle of
dependence of the axes orientations on the rock type (competent vs. Θ = 45°. These constraints are for example used in the P–B–T axes
non-competent). method (Turner, 1953) and in the numerical dynamic method (Spang,
1972).
3.3. Choice of method: influence of reactivation vs. newly formed faults Reactivation of pre-existing planes does not follow the Mohr–
Coulomb fracture criterion, which implies that the angle between
One prerequisite for this last separation step is the choice of a maximum principal stress axis and fault plane is not fixed, and oblique
suitable analysis method for each data set. The most important factor slip is a common feature. The maximum principal stress axis might lie
for this choice is the existence or non-existence of older planes which anywhere in the compression quadrant limited by the fault plane
might be reactivated. The formation of new faults in a homogeneous and the plane perpendicular to the slip direction (e.g. Angelier and
rock body (with no or only few pre-existing planes) appears to follow Mechler, 1977; McKenzie, 1969). The basic assumption used in the
the Mohr–Coulomb fracture criterion, with the maximum and analysis of reactivated faults is that slip occurs in the direction of
minimum principal stress axes lying in the plane containing the maximum shear stress along the fault plane. Analytical methods try to
fault plane normal and the slip vector (plane of movement; Arthaud, find best-fit reduced stress tensors (orientation of the three principal
1969). The angle Θ between the maximum principal stress axis and stress axes and value of stress ratio R) with smallest misfit between
the fault plane is fixed by rock properties, i.e. by the angle of internal observed and predicted fault–slip geometry like the direct inversion
friction φ (with Θ = 45 − φ / 2). Experimentally obtained values for Θ method (e.g. Angelier and Goguel, 1979) or the grid search method
range between 17° and 40° (e.g. Byerlee, 1968; Hubbert, 1951; Jaeger (e.g. Gephart, 1990). However, for these methods at least four
and Cook, 1979), so that an angle of Θ = 30° seems to be a reasonable differently oriented fault planes are necessary to unequivocally define
approximation. Similar geometrical relationships are valid for infini- the reduced stress tensor. Accordingly, fault sets with only two
B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41 33

Fig. 4. Separation of an artificial 2D-dataset into subsets using an automated procedure. (a) Original dataset with two clusters and calculated mean M1 lying between the clusters.
(b) Subset 1 obtained by minimising the deviation from M1. (c) Subsets 2 and 3 with means M2 and M3 obtained from the remaining data after separation of subset 1. Note that the
automated separation leads to three subsets being in disagreement with the original dataset with the two clusters. The largest subset 1 has a mean M1 with no relation to the original
data. Similar erroneous results might be obtained in fault–slip analysis when separation is purely done by minimising the misfit angle between theoretical and measured slip
direction. The only difference is that fault–slip analysis is carried out in 4D-space.

dominating orientations, as is typical for conjugate faults, give no Angelier, 1994; Meschede and Decker, 1993; Ramsay and Lisle, 2000;
reasonable result with these methods. Sperner, 1996; Twiss and Unruh, 1998) nor a listing of all available
software, but we want to point out the differences in faulting pro-
3.4. Comparison of methods: influence of different fault patterns and cesses (see previous section) and the error sources related to these
stress regimes differences. We thus compare for different fault patterns the results of
the numerical dynamic analysis (Spang, 1972; computer program NDA
In this paper, our intention is not a comprehensive discussion of all of Sperner and Ratschbacher, 1994) and the direct inversion method
available fault–slip analysis methods (overviews are presented e.g. in (Angelier and Goguel, 1979; computer program INVERS of Sperner

Fig. 5. Separation of a fault–slip data set (b157) into two subsets (b157a and b157b). Data are plotted in equal area, lower hemisphere projections using the tangent–lineation
method (slip direction of the hanging wall block is indicated by an arrow at the location of the fault plane normal). Obviously incompatible slickensides that have similar fault
orientations (small circles = poles) but different slip directions (arrows; for examples see encircled data) are separated into different data sets. The calculated principal stress axes
for the subsets are plotted as well. Data are listed in Appendix A.
34 B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41

et al., 1993) as two typical analysis methods for different faulting determined by any of the methods (although it is calculated by
scenarios comprising Mohr–Coulomb fracturing and reactivation of the software) and thus is useless for the interpretation.
pre-existing planes, respectively. (b) Conjugate fault sets (Fig. 6b): Conjugate sets are typical for
The numerical dynamic analysis was originally developed for the newly formed faults in homogeneous rocks with no or only a
analysis of calcite twins, but can also be applied to fault–slip data few pre-existing planes. The direct inversion method again fails
because twinning in calcite is analogous to slip along a fault (Larroque for the same reasons as discussed in (a), and the calculated
and Laurent, 1988; Shelley, 1992). The main difference, the fixed maximum and minimum principal stress axes are parallel to
orientation of the glide plane and of the glide direction as determined one of the fault sets, so that again no shear stress would build
by the crystal lattice, is compensated by the random orientation of the up along this fault set. The numerical dynamic analysis gives
crystals in an aggregate prior to deformation. In the numerical the expected result for conjugate fault sets with the interme-
dynamic analysis an arbitrary shear stress magnitude of 1 along each diate principal stress axis lying parallel to the intersection line
fault plane and a fixed angle between maximum principal stress axis of the two fault sets, and the maximum and minimum axes
and fault plane is assumed. The resulting stress tensors for the single bisecting them (the angle between maximum axes and fault
faults are added up to the bulk stress tensor defining the orientation planes is again assumed to be 30°). Again, the stress ratio R is
and relative values of the principal axes. In contrast, the direct inver- not defined for both methods (as long as the third dimension is
sion method calculates the reduced stress tensor in a single step from not involved into the movements along the faults, the
the entirety of the data by applying a least-square minimization of the equations for the stress tensor are underdetermined and R
shear stress component normal to the observed slip. For both methods can adopt any value).
the angles between theoretical and measured slip directions are (c) Axial extension (=constriction; Fig. 6c): Axial extension can
calculated by the software employed here; the mean of these angles easily be recognized in the tangent-lineation plot as the slip
defines the fluctuation F. Additionally, fault planes with slip sense arrows point away from one point, the minimum principal
opposite to the predicted one are labelled ‘nev’ (‘negative expected stress axis, and towards the plane containing the two other
value’; see Section 3.5). Data examples come from the Western principal axes. The latter axes are similar in magnitude, so that
Carpathians (Sperner et al., 2002); results are listed in Table 2. the stress ratio R will be near 1.0 for pure axial tension. Our
example includes a large number of conjugate normal faults, so
(a) Dominance of one fault plane orientation (Fig. 6a): This kind of
that the R value is shifted towards the plane stress configura-
data set might be the result of a preferred outcrop orientation
tion (R = 0.5) and thus shows avalue lower than the expected
(e.g. outcrop wall parallel to one set of a conjugate fault set) or
one (R = 0.6 instead of 1.0). The conjugate faults are also re-
of pre-existing planes in a preferred orientation. The direct
sponsible for the reasonable result revealed by the numerical
inversion method fails in this case, because it needs at least four
dynamic analysis. On the other hand the oblique slip data give
independent faults, i.e. four fault planes orientated differently
the necessary constraints for the direct inversion method, so
and with different slip directions. The calculated maximum
that in this case both methods reveal reasonable and almost
principal stress axis lies parallel to the fault planes, which is in
identical results.
accordance with minimum shear stress normal to slip, but not
(d) Axial contraction (= flattening; Fig. 6d): Axial contraction
with the mechanical requirement of a certain amount of shear
displays in the tangent–lineation plot a pattern similar to that
stress necessary for the initiation of slip even on pre-existing
for axial extension, but with opposite signs: the slip arrows
planes. The numerical dynamic analysis gives reasonable results
point towards one point, the maximum principal stress axis,
with the intermediate principal stress axis parallel to the fault
and away from the plane containing the other two principal
planes and the maximum and minimum axes at a distinct angle
axes. The latter axes are similar in magnitude, so that the stress
to the fault planes. A reasonable value for the angle between
ratio R will be near 0.0 for pure axial compression. In our
compression axis and fault plane is 30° for newly formed faults
example results from both methods give low R values (0.1 and
(derived from the angle of internal friction). For the case of
0.2), identical orientations for the maximum principal stress
reactivation this angle is not fixed and can vary for each fault
axes, and varying orientations for the other principal stress
plane, so that no statement about the orientation of the principal
axes. The direct inversion method reveals the ‘better’ result as it
stress axes is possible. Thus, analysis of such data sets only makes
gives a lower fluctuation (i.e. the arithmetic mean of the angle
sense if reactivation can be excluded. With only one set of
between the measured and the theoretical slip directions).
subparallel faults present, the stress ratio R cannot be reasonably
(e) Triaxial deformation (all three principal infinitesimal strain axes
are nonzero and different from each other in magnitude): We
don't have a typical example of this kind of deformation
(maybe rare in reality), but what can be expected is a scattering
of fault orientations and a high percentage of oblique slip data,
Table 2 both caused by the reactivation of planes with varying orien-
Parameters of the reduced stress tensors plotted in Fig. 6. σ1–3, orientation of principal tations. This makes the numerical dynamic analysis useless for
stress axes (σ1 ≥ σ2 ≥ σ3); R, stress ratio (σ2 − σ3)/(σ1 − σ3); F, fluctuation (arithmetic analysing this type of deformation. In contrast, the direct
mean of the angles between measured and theoretical slip directions); nev, “negative inversion method will reveal reasonable results.
expected value” (i.e. wrong shear sense); n, number of data in subset; N, total number
of data in outcrop; Q, quality of result (A, best quality, D, worst). More reliable results
(f) Non-coaxial deformation has been reported by Schrader (1991,
are printed in bold (see grey symbols in data plots of Fig. 6). 1995). Twiss and Unruh (1998) attributed this type of defor-
mation to a combination of local rigid block rotations with the
File (method) σ1 σ2 σ3 R F Nev n (N) Q
overall strain of a shear zone. They express, however, the need
b177 (NDA) 217 12 360 75 125 09 0.6 9° 0 11 (11) C for the presence of a multitude of differently oriented, pre-
b177 (INVERS) 195 08 096 47 292 41 0.02 3° 2 11 (11) C
existing weakness planes; a condition that is not generally
b156a (NDA) 040 20 198 69 307 08 0.5 9° 0 28 (39) A
b156a (INVERS) 001 10 269 08 141 77 0.2 8° 13 28 (39) A satisfied in weakly deformed rocks which are usually analysed
b170a (NDA) 322 63 122 26 216 08 0.6 13° 0 22 (71) C by fault–slip analysis. Nevertheless, caution is appropriate
b170a (INVERS) 296 77 118 13 028 00 0.6 11° 0 22 (71) C when applying most of the available methods for fault–slip
b169a (NDA) 190 14 092 30 302 57 0.1 10° 0 19 (22) D analysis to strongly deformed fault zones. There, the calculated
b169a (INVERS) 197 18 290 09 047 70 0.2 6° 0 19 (22) D
principal stress and strain axes may deviate considerably in
B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41 35

Fig. 6. Fault–slip analysis for different fault patterns analysed by means of the numerical dynamic analysis (Spang, 1972; computer program NDA of Sperner and Ratschbacher, 1994)
and the direct inversion method (Angelier and Goguel, 1979; computer program INVERS of Sperner et al., 1993). (a) Dominance of one fault plane orientation. (b) Conjugate strike–
slip faults. (c) Axial extension. (d) Axial contraction. (For legend see Fig. 5). For the principal stress axes the more suitable results are plotted in grey symbols, while the others are
coloured white. Histograms on the right side show the distribution of angles between measured and theoretical slip orientation for both methods. Results are listed in Table 2, data
and errors in Appendix B.

orientation from the actual ones (Twiss and Unruh, 1998). The tensor. Computer programs NDA and INVERS, for example, produce
principal strain axes even might vary with time as simulated by a data list indicating for each datum the specific error (= angle
Twiss et al. (1991) taking into account not only the symmetric between measured and theoretical slip orientation; Appendices A, B)
(non-rotational), but also the rotational part of the velocity and whether the slip sense fits. A non-fitting slip sense is in these
gradient tensor. programs marked by ‘nev’, the abbreviation for ‘negative
expected value’ with ‘negative’ meaning an opposite slip sense
3.5. Check of results compared with the theoretically determined one. Well-separated
data sets contain not even one ‘nev’-fault and the angle between
Fault–slip analysis always includes a check of the compatibility measured and theoretical slip direction does not exceed 30° for
of the slip sense on the fault planes with the calculated stress any fault.
36 B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41

Table 3 The average value for all structures used in the analysis gives
Quality ranking scheme for fault–slip data (Sperner et al., 2003; for further explanation
the “data-type number” for Table 3. For pure fault–slip data
see Section 3.6).
sets this number (=1.00) has no relevance.
Sub- Number Percentage Confidence number Fluctuation [°] Data
quality of data of data (slip-sense determ.) (average misfit) type
(f) Diversity of orientation of fault planes and of lineations. Fault–slip
a ≥25 ≥60 ≥0.70 ≤9 ≥0.90 analysis is based on an equation system with four unknowns
b ≥15 ≥45 ≥0.55 ≤12 ≥0.75 (e.g. Angelier and Goguel, 1979) due to the search for the
c ≥10 ≥30 ≥0.40 ≤15 ≥0.50
d ≥6 ≥15 ≥0.25 ≤18 ≥0.25
orientation of the three stress axes and for the stress ratio. Thus at
e b6 b 15 b0.25 N 18 b0.25 least four differently oriented fault planes are necessary to yield a
unique solution. Unbalanced data sets with fault planes and
The overall quality is defined by the lowest sub-quality of any of these criteria.
lineations with similar orientations will not give clear results for
3.6. Quality of result both the orientation of the stress axes and the stress ratio. Data
sets including only two fault plane orientations (conjugate fault
The quality of the results of fault–slip analysis is dependent on sets) give better, but still not optimum restrictions concerning
several different criteria such as the number of data and the average the stress axes, while the stress ratio is still undefined. Thus, the
misfit between measured and theoretical slip direction. In order to diversity of the measured data ensures the non-ambiguity of the
assure a better comparability of the data sets' quality, we recommend calculated result. A measure of the diversity is given by the
the use of the quality ranking scheme of the World Stress Map (WSM) normalised length of the mean vector of the fault plane poles and
project (Sperner et al., 2003) which we describe below and which is of the mean slip line vectors. This normalised length is close to 1
available at http://www.world-stress-map.org. for parallel planes or parallel lineations and could be used as an
The following criteria are used to define “sub-qualities” ranging additional quality criterion like it is done in the software package
from a to e (Table 3). The overall quality A to E (A — best quality, E — TENSOR of Delvaux and Sperner (2003). We plead for a careful
worst) of the analysis is given by its lowest sub-quality. handling of fault–slip data concerning this problem.

(a) Number of data. The number of data used for the calculation of Note that a data set must fulfil all requirements, i.e. a quality A data
the stress axes. set must comprise at least 25 fault–slip measurements, which
(b) Percentage of data. The percentage of data used in a (sub)set relative to represent at least 60% of all measured data in this outcrop and have
the total number of data measured in an outcrop. With more than one a confidence number larger than 0.70 and a fluctuation smaller than
event recorded in an outcrop, the quality will (in most cases) decrease 9°. Examples are given in Table 2. For the WSM data base, data with
reflecting the difficulty to correctly separate the data into subsets. unknown confidence level are treated like having the lowest
(c) Confidence number. The reliability of the results of fault–slip analysis confidence level 4, i.e. the confidence number is 0.25.
is very sensitive to the reliability of slip-sense determination.
Hardcastle (1989) introduced four confidence levels (1–4) for slip- 4. Steps 3 and 4: Age assignment and derivation of regional patterns
sense determination which are assigned to the single faults during
field measurements. These confidence levels are assigned different Fault–slip data are usually collected with the purpose of deriving a
weights in the following way: regional deformation pattern. This requires that the fault sets from
several locations have to be grouped according to the age or the age
1, absolutely certain: 1.00 span during which their faults were active. This, in turn, demands
2, certain: 0.75 dating of the individual fault sets. Theoretically, dating of fault sets
3, uncertain: 0.50
and derivation of a regional pattern should be two separated steps in
4, very uncertain: 0.25
fault–slip analysis. In practice, however, they are frequently com-
bined; therefore, they will be discussed in one section here.
The “confidence number” is defined as the average of these values
for all data used in the final analysis. If no confidence levels were 4.1. Direct dating
assigned during field measurements the overall quality assigned
due to the “rest” of the criteria is taken and downgraded by one Direct dating of the age of individual faults is only possible in rare
class (e.g. from quality B to C). cases. Even though radiometric methods for dating of mineralisation
on faults (e.g. Eyal et al., 1992) or of fault gouges (e.g. Hataya et al.,
(d) Fluctuation. The fluctuation (in some publications called “misfit 1997; Kralik et al., 1987) exist, they can only seldom be applied. In
angle”) is defined as the arithmetic mean of the angles between many cases, a fault set affects a rock unit that can be dated (usually
the measured and the theoretical slip directions (the theoret- sediments), the rock age thus providing a lower age limit. In some
ical slip direction is assumed to be parallel to the direction of rather uncommon cases, these faults terminate against unconformi-
maximum shear stress along the fault plane). ties, such that the age of the overlying rocks yields an upper age limit
(e) Type of data. Information from other structures like tension which effectively brackets the time span of faulting activity.
joints can give additional constraints on the calculated stress
tensor. The computer program TENSOR written by Delvaux 4.2. Faults and folds
(Delvaux et al., 1997; Delvaux and Sperner, 2003) and available
at http://users.skynet.be/damien.delvaux/Tensor/tensor-index. In some cases, the interrelation of faults with folds is visible in
html includes these data into the analysis by using different outcrops, e.g. faults of one set being folded (i.e. being older than the
weights for the different types of structures: fold), whereas another fault set cross-cuts folds, thus evidently
postdating the folds. If the folds can be dated, e.g. by knowing the age
Slickenside: 1.00 of the youngest rocks involved in folding and/or the age of a
Tension/compression joint: 0.50 discordantly overlying unit, this provides age constraints on fault
Two conjugate planes: 0.50 formation and fault activity. Yamaji et al. (2005) recently implemen-
Movement plane + tension joint: 0.50 ted a long used method into a software package, the back rotation of
Shear joint: 0.25
fold limbs by which fault data from both limbs can be analysed. The
B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41 37

method also gives information on the relative ages of folding and direction of the maximum principal stress axis which is then assumed to
faulting, i.e. whether faulting occurred before, during or after folding, portray the same deformation event as parallel-oriented maximum
but it requires a separation of the data set into subsets for each fold principal stress axes derived from faults in neighbouring exposures (e.g.
flank that merge after rotation into one consistent data set. Huismans et al., 1997). This assumption may in many cases be correct.
A much more frequent but intricate situation is, however, that a However, the bearing of folds is often not normal to regional compression
fault-containing outcrop is situated on a single, uniformly dipping (maximum regional principal stress axis) but is frequently deflected from
limb of a large fold such that a clear pre- or post-folding relationship is such an orientation due to, e.g., mechanical inhomogeneities in the
not evident. Nevertheless, pre-folding faults may be present which deformed rocks such as variations in the effectivity of an underlying
became rotated during folding. A typical assumption is that the brittle detachment or inherited structural grain (previous folds or faults).
fault sets were formed close to the earth's surface where, according to
the theory of Anderson (1951), all principal stress axes should be 4.4. ‘Correlation’ of a fault set to age-constrained fault sets from other
either horizontal or vertical. A usual line of reasoning is then as follows: locations

(i) If all calculated principal stress axes of a fault set are roughly Correlation of fault sets that yielded comparably oriented principal
horizontal and/or vertical, an activity of the faults after folding is stress axes is a similar, if not still more common method to derive an
assumed. age for a fault set. The assumption underlying this procedure is again
(ii) If all or some of the principal stress axes are neither horizontal that stress is homogeneous within the region of interest, which is not
nor vertical, a pre-folding activity of the faults is suspected and a necessarily the case, as argued above. Moreover, the principal stress
back-rotation might be appropriate. If all principal axes then are axes of syngenetic faults have been shown to exhibit a systematic
change of orientation in areas where differently oriented, but syn-
roughly horizontal and/or vertical, a pre-folding activity of the
chronously active fault zones intersect (Dupin et al., 1993; Merle et al.,
faults will be assumed.
1993; Pollard et al., 1993; Zweigel et al., 1998). Further complications
Calculated principal axes deviating from horizontal or vertical may be due to block rotation leading to passive rotation of previously
positions, however, also may be the consequence of non-Andersonian active faults (e.g. Márton and Fodor, 1995). In conclusion, the as-
stress states at depth caused, for example, by lateral density variations sumption that principal stress axes from synchronously formed faults
or other changes in rock properties (e.g. competence contrast). Such in a certain region should be parallel must be treated with consid-
effects might influence the orientation of principal stress axes without erable caution.
being as obvious as in the case of oblique axes. Frequently, the outlined methods to derive ages for individual
fault sets by reference to dated features from other localities are used
4.3. Correlation of a fault set to a major fault, fault zone or fold of known extensively, because faulting is in many cases rather scarce in syn-
age deformational sediments, but is abundant in neighbouring, older,
more lithified rocks. Implicitly, the allegedly contemporaneously
In the frequent cases that an adequate (i.e. precise enough) age formed fault sets can then NOT be conclusively used to derive a regional
assignment by outcrop data is not possible, more speculative chains of stress pattern for the time step of interest. This is because the age
reasoning have to be employed, such as correlation of a fault set to a assignment had been done under the assumption that fault sets with
major (mapped) structure of known age or to one or several dated the same orientation of principal stress axes have the same age. The
fault sets from other locations. The former is usually done if a major conclusion that the stress pattern proves a homogeneous deformation
fault is situated close to the analysed outcrop, i.e. up to a few km (!) in of the study area would then be a circular argumentation.
distance, and if the fault set contains a major group of faults that are
either parallel to the fault zone itself or parallel to orientations in 4.5. Comparison of regional patterns
which subordinate faults linked to a major fault zone occur (e.g. Riedel
faults, Fig. 7). As is evident from Fig. 7, there exist subordinate faults The orientations of principal stress axes from fault–slip analysis
in various orientations, and if reasonable uncertainty brackets (e.g. are typically plotted on maps as a means to identify regional stress
± 20°) are used, a ‘correlation’ of a fault set to a major fault is nearly patterns. It is for example common to plot the horizontal projections
always possible. A minimum demand in such a process would be that of the maximum principal stress axes for the data sets that indicate
the slip sense along the fault group accords with the slip along the strike–slip or reverse faulting (maximum principal stress axis rather
subordinate fault as predicted from theory. horizontal than vertical). Lund and Townend (2007) pointed out that
Such a ‘correlation’ is, however, not at all conclusive. Several
assumptions underlie the outlined age assignment procedure, e.g. that
deformation within a rock body (e.g. a nappe sheet) is essentially a
down-scaled mirror of the deformation in the fault zone at the margin
of that rock body (e.g. a major thrust), that stress is homogeneous within
the deformed rock body, and that deformation within the rock body
must be contemporaneous with deformation at its margins. All these
three listed assumptions are in many cases not valid, e.g. because rock
mechanical properties are generally not homogeneous within any km-
scale rock body (e.g. Heidbach et al., 2007; Müller et al., 1997).
Moreover, stress trajectories change orientation systematically when
approaching a major fault zone (e.g. rotated stress field near the San
Andreas fault; Zoback et al., 1987) such that the orientation of the
principal stress axes in the surrounding of a fault zone may be far from
being parallel to those corresponding to the fault zone itself.
Similarly, fault sets are in some cases dated by ‘correlation’ from an
exposure to other structures of known or inferred age occurring within
some distance, such as mapped or exposed folds. In the case of folds, the Fig. 7. Fault zone and related brittle structures. σ1,3, principal stress axes; M, main fault
direction normal to the bearing of the fold axis is taken to be the local zone; P, P shear; R, Riedel shear ; R', conjugate Riedel shear; T, tension fracture.
38 B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41

the projected axis can deviate by tens of degrees from the orientation (e) Documentation of relative ages (e.g. crosscutting relationships,
of the true maximum horizontal stress (SH). However, for such cases, curved slip indicators).
for which it would be reasonable to include SH from a data set into a (f) Documentation of absolute age indications (e.g. sealing sedi-
regional evaluation (strike–slip regime, or reverse regime with non- ments on top of faults).
oblate stress) the deviations are typically below 10° and thus within (g) Documentation of type of fault gouge and mineralization
the range of methodological uncertainty. For specific cases with large (might reflect different formation conditions, e.g. different
deviation, the correction method developed by Lund and Townend temperatures/depths)
(2007) may be used to reduce the error introduced by projection into (g) Documentation of indicators for reactivation vs. new formation
the horizontal plane. of faults.
Smoothing of orientation data can give an overview about the (h) Documentation of additional features (see Table 1).
general orientation of the stress field, but also bears the risk that the
data are overinterpreted in the way that trajectories are intra- or
5.2. (2) Data analysis
extrapolated to regions with no or only few data points. Thus, the
smoothing parameters, like the number of next neighbours used in
the calculation, have to be chosen with care (Müller et al., 2003). (a) Check the data for consistency (slip lineation parallel to asso-
ciated fault?).
(b) Do the first rough separation due to field criteria or, if not
4.6. Conclusion
available, due to compatibility criteria(same orientation of fault
planes, but different orientations of lineations indicate different
We want to emphasise that reasoning chains as presented above
events; BUT: other options are e.g. block rotations).
may still be valid in several cases, e.g. the interpretation that fault sets
(c) Check the remaining data for compatibility with the already
with parallel principal stress axes in a region are syngenetic. However,
separated data subsets.
the chains underlying the critical step of age assignment are often
(d) Choose an appropriate method depending on whether the faults
not clearly stated, and the reader has thus frequently no possibility to
are newly formed (mainly conjugate faults) or reactivated
evaluate the consistency and the regional meaning of the data. Besides,
(oblique slip prevailing). Different subsets from a single outcrop
if, e.g., parallelism of principal stress axes was used as an assumption
might have different origin and require different treatment.
during age assignment to fault sets, authors seem not always to be
(e) Check the compatibility of measured and theoretical slip direc-
aware of the limitations imposed on the regional interpretation of their
tions, i.e. slip sense (‘nev’) and misfit angle (b30°).
data or at least they do not mention these limitations in their discussion.
(f) Determine the data quality according to WSM quality ranking
scheme.
5. Our plea: documentation! (g) Plot the data subsets as tangent–lineation plots including the
orientations of the principal stress axes. Data not used in any of
Evaluation of the quality of fault–slip data, of their analysis and their the subsets are presented in a separate plot, if they account for
interpretation is in many publications hindered by the lack of docu- more than 20% of the whole data set.
mentation. Usually, only data plots are presented, in most cases after
separation was carried out. A complete list of the raw, unseparated data
5.3. (3) Data interpretation
is seldom given. Thus, the reader has no possibility to test other (new)
methods or software with the published data and to compare the
(a) Avoid over-interpretation: for some fault patterns even the
results. Different strategies in separation might lead to different data
reduced stress tensor can not be determined unequivocally
subsets and thus different results, which can only be checked if the
(e.g. no stress ratio definable for purely conjugate faults).
complete data sets are published, including the data not used for the
(b) Sound age assignments. Best are absolute movement ages
current analysis because of incompatibility. Additional information like
(dated slickenfibres or fault gouges) and lower/upper age
age relations between faults, outcrop conditions, age of rocks, etc. are
limits given by the age of the rock containing the faults/sealing
further details to be documented together with the raw data. In the days
the faults on top.
of Internet and the option to publish additional material online, offered
(c) Interrelations with folds might give indications for pre- or post-
by most of the journals, lack of space is no longer an excuse for incom-
folding deformation, but folds tend to rotate around vertical
plete data documentation.
axes.
Besides of the documentation of the raw data themselves, the
(d) Grouping of the results according to the orientation of the
strategies for data separation as well as the arguments for the choice
principal stress axes only makes sense when local or regional
of a method are important information in any publication using fault–
effects (e.g. rotating blocks, fan-shaped deformation patterns in
slip analysis. Age assignments need to be discussed in detail, so that
collision zones) can be excluded.
their reliability can be assessed.
We propose the following steps for a sophisticated fault–slip
analysis: 6. Meta-methodological implications

In our opinion, fault–slip analysis is now in a situation that may be


5.1. (1) Data acquisition
rather typical for many scientific methods when reaching a certain
level of maturity and acceptance in the community. It is now a
(a) Accurate field measurements (only faults with known slip blooming method, being widely used to understand the kinematic
sense). and/or dynamic evolution in many tectonic settings. For some years
(b) In case of large (undulating) fault planes: use average of several now, fault–slip analysis has become part of the standard curriculum
measurements. as expressed, e.g., by the fact of being treated in structural geology
(c) Quality assignment (confidence level) for slip–sense textbooks (e.g. Meschede, 1994; Ramsay and Lisle, 2000; Twiss and
determination. Moores, 1992). Currently published papers on fault–slip analysis
(d) No (unconscious) selection of data (e.g. by measuring mainly present in some cases modifications or refinements of the method
the more obvious faults subparallel to the outcrop wall, while (e.g. Delvaux and Sperner, 2003; Liesa and Lisle, 2004; Lisle and
faults oblique to the wall are missed). Srivastava, 2004; Orife and Lisle, 2003; Xu, 2004; Yamaji, 2000).
B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41 39

However, most papers simply employ the method without question- from calculated stress axes determined for b157b with program NDA:
ing it, and the authors often no longer see the need to explain in any σ1 = 197 26, σ2 = 037 63, σ3 = 291 08, R = 0.504, and for b157a with
detail the method and their way of handling it. program INVERS: σ1 = 190 82, σ2 = 334 07, σ3 = 064 05, R = 0.087.
The acceptance of fault–slip analysis as a standard method results
in its increased use by persons who were not involved in method
development and who frequently apply it rather uncritically just No. Plane Glide Sense Conf. Error nev?
because they consider the method to be well-established. Which
b157b
method (i.e. which analytical procedure, which software package etc.) 1 12 60 305 34 − 2 8
these users follow, is often more a matter of chance (which software is 2 327 72 250 34 − 3 10
easily available, which paper is easy at hand) than the outcome of a 3 344 84 260 45 − 2 19
conscious decision. This common behaviour is partly caused by the 4 357 72 283 40 − 1 14
5 329 62 245 11 − 3 11
fact that the practical application of many scientific methods is — in 6 305 81 219 21 − 2 11
contrast to their theoretical base — not documented in detail in 7 316 83 45 11 + 2 32
scientific papers nor in textbooks. Many specific methods represent 8 63 50 347 16 + 1 29
too narrow a ‘niche’ in their discipline to be thought of to merit 9 146 76 228 28 + 1 13
10 57 50 344 20 + 1 27
detailed procedural descriptions. (A rare exception for fault–slip
11 52 47 351 28 + 1 31
analysis is its treatment in Meschede and Decker (1993), which is 12 84 87 173 26 − 1 15
however not readily understandable for non-German speakers.) New 13 149 87 239 7 + 1 14
practitioners of such methods can thus only learn them properly by 14 143 74 231 8 + 1 2
direct instruction from an experienced senior colleague. 15 139 89 229 19 + 1 3
16 293 72 213 29 − 1 19
In a mature stage of a practical scientific method, critical papers 17 319 77 238 33 − 1 4
that point to the potential for mistakes, mis-interpretations or mis- 18 143 85 231 20 + 1 0 nev
use of the methodology are often lacking. This is in striking contrast to 19 326 80 237 7 − 1 18
the general view of science practitioners who see science as a critical 20 139 83 226 23 + 1 4
21 299 65 226 32 − 1 14
enterprise. In our opinion, such criticism must cover not only theories
22 53 89 142 38 − 1 21
(e.g. about faulting processes, or about the tectonic evolution of an 23 105 86 16 16 − 3 0
area) but also the technical methods that are employed to generate 24 142 89 231 56 + 2 29
the data necessary to build or test the theories. We would thus like to 25 142 79 211 62 + 2 38
call for both: better description of method application in the public 26 290 80 217 58 − 2 6
27 303 74 230 46 − 1 2
domain and more methodological criticism. 28 348 85 262 42 − 1 14
Similarly, lacking documentation of primary data and of their
analysis in papers applying ‘standard methods’ inhibits critical assess- b157a
ment of the results. Commonly, belief in the authors' integrity and 1 8 55 38 51 − 2 15
2 307 84 27 61 − 1 22
technical capacity is required, and no explicit test of their work is
3 321 67 338 66 − 1 16
possible. In the case of fault–slip analysis, documentation of measured 4 52 54 65 53 − 2 0
data is easily possible and modern communication facilities can make 5 6 61 25 59 − 3 6
these data widely accessible. (Note that it is still necessary to believe 6 13 82 41 81 − 1 4
in the quality of the authors' measurements and observations in the 7 105 86 190 47 − 4 12
8 348 75 327 74 − 3 5
outcrop, and testing of the primary data would entail substantial 9 116 79 193 51 − 1 19
effort in terms of cost and time.) Our plea is thus to improve both 10 132 81 211 50 − 1 13
methodological documentation (in textbooks and in papers applying 11 183 31 223 25 − 1 27
the method) and data documentation. We are convinced that this will 12 233 82 241 82 − 4 2
13 244 88 158 66 + 3 2
be an important component towards making structural geology a
14 268 88 345 81 + 4 5
really quantitative science. 15 228 55 258 51 − 2 20
16 217 43 217 43 − 4 1
17 289 63 249 56 − 4 6
Acknowledgments 18 297 61 249 50 − 4 14
19 271 89 353 80 + 4 5
20 83 85 162 66 − 4 10
Our special thanks go to our former PhD supervisor Lothar
21 32 54 23 54 − 1 12
Ratschbacher (now Freiberg University) for introducing us to the 22 26 87 46 87 − 2 7
methods of fault–slip analysis and for his permanent demand for high 23 39 59 55 58 − 1 1
work standards. Funding of much of our fault–slip analysis related work, 24 38 74 54 73 − 1 4
which we mainly undertook at the University of Tübingen (Germany), 25 22 50 41 48 − 2 6
26 323 88 240 76 − 1 0
was by the Deutsche Forschungsgemeinschaft for BS (Fr 610/9) and by a
27 29 60 64 55 − 4 12
doctoral research grant (GraFöG) of the government of Baden- 28 37 81 103 68 − 1 6
Württemberg to PZ. We thank Bernard Célérier, Damien Delvaux, 29 46 73 78 70 − 1 2
Richard Lisle and two anonymous reviewers for their comments on the 30 52 47 76 44 − 1 7
31 218 30 211 30 − 2 2
actual as well as on an earlier version of the manuscript.
32 18 68 5 67 − 2 11
33 63 42 73 42 − 1 2
Appendix A 34 73 65 95 63 − 3 4
35 254 23 216 18 − 3 14
36 248 51 220 47 − 4 10
Data example used for the separation in Section 3.2. (see also 37 241 22 223 21 − 1 3
Fig. 5). ‘Plane’, dip azimuth and plunge; ‘Glide’, bearing and plunge; 38 213 29 229 28 − 3 14
‘Sense’, slip sense with ‘−‘ for normal and ‘+’ for reverse faulting; 39 16 86 65 83 − 1 5
‘Conf’, confidence level (see Fig. 5); ‘Error’, angle between measured 40 28 61 51 59 − 1 5
41 217 34 217 34 − 1 2
and calculated slip orientation; ‘nev?’, opposite slip sense as predicted
40 B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41

Appendix B Appendix B (continued)


(continued)
No. Plane Glide Sense Conf. Error nev? Error nev?
Data used for the comparison of methods in Section 3.4. (see also b170a NDA INVERS
Fig. 6). ‘Plane’, dip azimuth and plunge; ‘Glide’, bearing and plunge; 21 234 62 166 36 − 3 31 15
‘Sense’, slip sense with ‘−‘ for normal and ‘+’ for reverse faulting; ‘Conf’, 22 50 28 29 26 − 2 9 6
confidence level (see Fig. 5); ‘Error’, angle between measured and
b169a
calculated slip orientation; ‘nev?’, opposite slip sense as predicted from 1 109 64 23 8 + 3 12 1
calculated stress axes determined for both methods ‘numerical dynamic 2 301 84 212 8 − 4 14 1
analysis’ (program NDA) and ‘direct inversion’ (program INVERS). 3 126 77 205 39 + 4 7 8
4 91 80 2 3 + 3 4 18
5 86 57 174 2 − 3 8 5
6 49 29 13 24 + 4 18 4
No. Plane Glide Sense Conf. Error nev? Error nev? 7 273 19 197 5 − 4 15 1
8 179 48 199 46 + 4 5 2
b177 NDA INVERS
9 203 38 211 38 + 3 8 12
1 99 82 10 5 + 2 23 3
10 137 61 209 29 + 4 4 9
2 268 82 357 4 − 2 15 9
11 229 64 180 54 + 4 17 6
3 105 88 194 17 − 1 3 1
12 303 66 32 3 + 4 11 2
4 283 84 196 23 + 1 8 2
13 177 35 154 33 + 3 32 18
5 85 89 175 7 − 2 8 0
14 258 50 192 27 + 3 3 10
6 100 84 188 15 − 1 5 4
15 305 49 24 12 + 3 7 7
7 114 85 203 9 − 1 13 1 nev
16 136 58 198 38 + 2 5 2
8 260 83 172 12 + 1 1 1
17 133 57 202 30 + 2 2 9
9 97 87 186 18 − 1 0 6
18 65 43 356 19 + 3 7 1
10 265 74 179 12 + 1 4 0
19 57 79 142 23 − 4 20 5
11 110 79 171 68 − 3 22 8 nev

b156a
1 179 83 93 29 − 4 19 4
2 106 84 18 21 + 4 4 7 nev References
3 257 86 345 29 − 3 9 21
4 86 89 356 20 + 3 1 8 Anderson, E.M., 1951. The Dynamics of Faulting. Oliver & Boyd, Edinburgh. (206 pp).
5 118 88 29 22 + 4 1 6 nev Angelier, J., 1994. Fault slip analysis and palaeostress reconstruction. In: Hancock, P.L.
6 113 83 26 21 + 4 4 10 nev (Ed.), Continental deformation. Pergamon Press, pp. 53–100.
7 180 86 94 43 − 1 32 7 Angelier, J., Goguel, J., 1979. Sur une methode simple de determination des axes princepaux
8 92 85 3 13 + 1 4 3 nev des contraintes pour une population de failles. Comptes Rendus Hebdomadaires des
9 100 89 10 19 + 1 0 2 nev Seances de l'Academie des Sciences, Serie D: Sciences Naturelles 288 (3), 307–310.
Angelier, J., Mechler, P., 1977. Sur une méthode graphique de recherche des constraintes
10 124 70 38 12 − 1 13 17
principales également utilisable en tectonique et en séismologie: la méthode des
11 117 87 27 6 + 1 6 9 nev
dièdres droits. Bulletin de la Societe Géologique de France 19 (6), 1309–1318.
12 262 89 352 1 − 2 19 1
Arthaud, F., 1969. Méthode de détermination graphique des directions de raccourcisse-
13 118 88 28 13 + 4 5 3 nev ment, d'allongement et intermédiaire d'une population de failles. Bulletin de la
14 106 83 19 21 + 3 5 8 nev Société Géologique de France 11 (5), 729–737.
15 189 83 279 2 + 3 12 5 nev Byerlee, J.D., 1968. Brittle–ductile transition in rocks. Journal of Geophysical Research
16 170 76 88 28 − 2 19 11 73 (14), 4741–4750.
17 157 78 71 17 − 2 5 10 Delvaux, D., Sperner, B., 2003. New aspects of tectonic stress inversion with reference to
18 268 84 357 13 − 2 7 11 the TENSOR program. Geological Society, London, Special Publications 212, 75–100.
19 259 87 348 15 − 1 5 10 Delvaux, D., Moeys, R., Stapel, G., Petit, C., Levi, K., Miroshnichenko, A., Ruzhich, V.,
20 162 80 73 4 − 1 5 0 San'kov, V., 1997. Paleostress reconstructions and geodynamics of the Baikal
21 58 86 329 16 + 1 14 0 region, Central Asia, Part 2. Cenozoic rifting. Tectonophysics 282 (1–4), 1–38.
22 13 88 103 7 + 3 11 22 nev Doblas, M., 1998. Slickenside kinematic indicators. Tectonophysics 295 (1–2), 187–197.
Dupin, J.M., Sassi, W., Angelier, J., 1993. Homogeneous stress hypothesis and actual fault
23 281 88 10 30 − 4 8 3 nev
slip: a distinct element analysis. Journal of Structural Geology 15 (8), 1033–1043.
24 324 81 46 42 + 1 26 7
Eyal, Y., Kaufman, A., Bar−Matthews, M., 1992. Use of 230Th/U ages of striated carnotites
25 74 89 344 27 + 1 5 16
for dating fault displacements. Geology 20 (9), 829–832.
26 61 87 333 31 + 1 0 17 Gephart, J.W., 1990. FMSI; a Fortran program for inverting fault/slickenside and
27 103 72 15 7 + 1 4 5 nev earthquake focal mechanism data to obtain the regional stress tensor. Computers &
28 94 74 11 23 + 1 12 10 nev Geosciences 16 (7), 953–989.
Hancock, P.L., 1985. Brittle microtectonics; principles and practice. Journal of Structural
b170a Geology 7 (3–4), 437–457.
1 315 44 20 22 − 1 6 24 Hardcastle, K.C., 1989. Possible paleostress tensor configurations derived from fault–
2 70 66 2 41 − 1 11 2 slip data in eastern Vermont and western New Hampshire. Tectonics 8 (2),
3 99 86 13 48 − 2 31 18 265–284.
4 51 53 11 46 − 2 3 1 Hataya, R., Tanaka, K., Miki, T., 1997. Studies on a new ESR signal (R signal) of fault
5 70 60 6 37 − 4 6 3 gouges for fault dating. Quaternary Science Reviews 16 (3–5), 477–481.
6 48 70 338 43 − 2 2 7 Heidbach, O., Reinecker, J., Tingay, M., Müller, B., Sperner, B., Fuchs, K., Wenzel, F., 2007.
Plate boundary forces are not enough: second- and third-order stress patterns
7 23 76 99 44 − 3 6 15
highlighted in the World Stress Map database. Tectonics 26, TC6014. doi:10.1029/
8 97 89 8 41 − 1 28 18
2007TC002133.
9 359 51 36 45 − 1 15 3 Hoeppener, R., 1955. Tektonik im Schiefergebirge; eine Einführung. Geologische Rundschau
10 202 79 275 57 − 2 1 3 44, 26–58.
11 180 69 240 52 − 2 10 8 Hubbert, M.K., 1951. Mechanical basis for certain familiar geologic structures. Geological
12 237 73 183 63 − 3 5 19 Society of America Bulletin 62 (4), 355–372.
13 206 76 272 59 − 2 2 8 Huismans, R.S., Bertotti, G., Ciulavu, D., Sanders, C.A.E., Cloetingh, S., Dinu, C., 1997.
14 222 58 176 48 − 3 31 14 Structural evolution of the Transylvanian Basin (Romania); a sedimentary basin in
15 191 64 264 32 − 1 19 22 the bend zone of the Carpathians. Tectonophysics 272 (2–4), 249–268.
16 191 70 264 39 − 2 10 11 Jaeger, J.C., Cook, N.G., 1979. Fundamentals of Rock Mechanics. Chapman & Hall, London.
17 350 88 80 2 − 2 26 4 (593 pp.).
18 179 86 268 17 − 1 15 1 Kralik, M., Klima, K., Riedmüller, G., 1987. Dating fault gouges. Nature 327 (6120),
19 179 86 264 47 − 4 11 17 315–317.
Larroque, J.M., Laurent, P., 1988. Evolution of the stress field pattern in the south of the
20 113 54 60 39 − 3 12 14
Rhine Graben from the Eocene to the present. Tectonophysics 148 (1–2), 41–58.
B. Sperner, P. Zweigel / Tectonophysics 482 (2010) 29–41 41

Liesa, C.L., Lisle, R.J., 2004. Reliability of methods to separate stress tensors from Shelley, D., 1992. Calcite twinning and determination of paleostress orientations; three
heterogeneous fault–slip data. Journal of Structural Geology 26 (3), 559–572. methods compared. Tectonophysics 206 (3–4), 193–201.
Lisle, R.J., Vandycke, S., 1996. Separation of multiple stress events by fault striation Spang, J.H., 1972. Numerical method for dynamic analysis of calcite twin lamellae.
analysis: an example from Variscan and younger structures at Ogmore, South Geological Society of America Bulletin 83 (2), 467–471.
Wales. Journal of the Geological Society 153 (6), 945–953. Sperner, B., 1996. Computer Programs for Kinematic Analysis of Brittle Deformation
Lisle, R.J., Srivastava, D.C., 2004. Test of the frictional reactivation theory for faults and Structures and the Tertiary Tectonic Evolution of the Western Carpathians
validity of fault–slip analysis. Geology 32 (7), 569–572. (Slovakia). Ph.D. Thesis, Tübingen University, 120 pp.
Lund, B., Townend, J., 2007. Calculating horizontal stress orientations with full or partial Sperner, B., Ratschbacher, L., 1994. A Turbo Pascal program package for graphical
knowledge of the tectonic stress tensor. Geophysical Journal International 170 (3), presentation and stress analysis of calcite deformation. Zeitschrift der Deutschen
1328–1335. Geologischen Gesellschaft 145, 414–423.
Márton, E., Fodor, L., 1995. Combination of palaeomagnetic and stress data: a case study Sperner, B., Ratschbacher, L., Ott, R., 1993. Fault-striae analysis; a Turbo Pascal program
from North Hungary. Tectonophysics 242 (1–2), 99–114. package for graphical presentation and reduced stress tensor calculation.
McKenzie, D.P., 1969. The relation between fault plane solutions for earthquakes and Computers & Geosciences 19 (9), 1361–1388.
the directions of the principal stresses. Bulletin of the Seismological Society of Sperner, B., Ratschbacher, L., Nemcok, M., 2002. Interplay between subduction retreat
America 59, 591–601. and lateral extrusion: tectonics of the Western Carpathians. Tectonics 21 (6), 1051.
Merle, O.R., H., D.G., Nickelsen, R.P., Gourlay, P.A., 1993. Relation of thin-skinned doi:10.1029/2001TC901028.
thrusting of Colorado Plateau strata in southwestern Utah to Cenozoic magmatism. Sperner, B., Müller, B., Heidbach, O., Delvaux, D., Reinecker, J., Fuchs, K., 2003. Tectonic
Geological Society of America Bulletin 105 (3), 387–398. stress in the Earth's crust: advances in the World Stress Map project. Geological
Meschede, M., 1994. Methoden der Strukturgeologie. Enke, Stuttgart. (168 pp). Society, London, Special Publications 212, 101–116.
Meschede, M., Decker, K., 1993. Störungsflächenanalyse am Nordrand der Ostalpen – Turner, F.J., 1953. Nature and dynamic interpretation of deformation lamellae in calcite
ein Vergleich numerischer und graphischer Methoden. Zeitschrift der Deutschen of three marbles. American Journal of Science 251, 276–298.
Geologischen Gesellschaft 144 (2), 419–433. Twiss, R.J., Gefell, M.J., 1990. Curved slickenfibers: a new brittle shear sense indicator
Müller, B., Wehrle, V., Zeyen, H., Fuchs, K., 1997. Short-scale variations of tectonic with application to a sheared serpentinite. Journal of Structural Geology 12 (4),
regimes in the western European stress province north of the Alps and Pyrenees. 471–481.
Tectonophysics 275 (1–3), 199–219. Twiss, R.J., Moores, E.M., 1992. Structural Geology. Freeman, New York. (532 pp).
Müller, B., Wehrle, V., Hettel, S., Sperner, B., Fuchs, K., 2003. A new method for Twiss, R.J., Unruh, J.R., 1998. Analysis of fault slip inversions; do they constrain stress or
smoothing orientated data and its application to stress data. Geological Society, strain rate? Journal of Geophysical Research 103 (6), 12,205–12,222.
London, Special Publications 209, 107–126. Twiss, R.J., Protzman, G.M., Hurst, S.D., 1991. Theory of slickenline patterns based on the
Orife, T., Lisle, R.J., 2003. Numerical processing of palaeostress results. Journal of velocity gradient tensor and microrotation. Tectonophysics 186 (3–4), 215–239.
Structural Geology 25 (6), 949–957. Xu, P., 2004. Determination of regional stress tensors from fault–slip data. Geophysical
Petit, J.P., 1987. Criteria for the sense of movement on fault surfaces in brittle rocks. Journal International 157 (3), 1316–1330.
Journal of Structural Geology 9 (5–6), 597–608. Yamaji, A., 2000. The multiple inverse method: a new technique to separate stresses
Pollard, D.D., Saltzer, S.D., Rubin, A.M., 1993. Stress inversion methods; are they based from heterogeneous fault–slip data. Journal of Structural Geology 22 (4), 441–452.
on faulty assumptions? Journal of Structural Geology 15 (8), 1045–1054. Yamaji, A., Tomita, S., Otsubo, M., 2005. Bedding tilt test for paleostress analysis. Journal
Ramsay, J.G., Lisle, R.J., 2000. The Techniques of Modern Structural Geology, Volume 3: of Structural Geology 27 (1), 161–170.
Applications of Continuum Mechanics in Structural Geology. Academic Press, Zoback, M.D., Zoback, M.L., Mount, V.S., Suppe, J., Eaton, J.P., Healy, J.H., Oppenheimer, D.H.,
London. (361 pp). Reasenberg, P.A., Jones, L.M., Raleigh, C.B., Wong, I.G., Scotti, O., Wentworth, C.M., 1987.
Schrader, F., 1991. Field-based analysis of faults, stylolites and extension gashes. New evidence on the state of stress of the San Andreas fault system. Science 238
Annales Tectonicae 5 (2), 91–101. (4830), 1105–1111.
Schrader, F., 1995. A new technique to determine deformation directions in multi- Zweigel, P., Ratschbacher, L., Frisch, W., 1998. Kinematics of an arcuate fold–thrust belt:
deformational fault assemblages, applied in the Cantabrian Mountains, Northern the southern Eastern Carpathians (Romania). Tectonophysics 297 (1–4), 177–207.
Spain. Zentralblatt für Geologie und Paläontologie, Teil 1 1994 (5/6), 447–458.

S-ar putea să vă placă și