Laurich2014 PDF
Laurich2014 PDF
Laurich2014 PDF
a r t i c l e i n f o a b s t r a c t
Article history: Slickensided shear surfaces are ubiquitous in many fault zones. However the internal structure, the
Received 27 March 2014 micromechanics and the evolution of these structures are not fully understood and the contributions of
Received in revised form crystal plasticity, grain-boundary sliding, microfracturing, solution-precipitation and mineral trans-
23 July 2014
formation under different conditions are subject of debate.
Accepted 24 July 2014
Available online 4 August 2014
We studied well- preserved core samples from the Main Fault, an up to 3 m wide zone of approxi-
mately 10 m offset in the Mont Terri Underground Research Laboratory (CH), a site to evaluate long-term
safety of radioactive waste disposal. The drill core breaks easily along many slickensided shear surfaces
Keywords:
Shear zone
indicating reverse slip, which form an anastomosing network connected by branch lines.
Slickenside Broad ion beam polishing and scanning electron microscopy shows that the slickensides are invariably
Scaly clay revealed by fracture of the drill core along a few mm thick shear zone, which acts as a crack guide for
Gouge fracturing the samples. In this zone, a complex set of processes is inferred, leading to extreme localization
Opalinus clay of strain, development of strong particle preferred orientation, the formation of nanoparticles, and local
Electron microscopy precipitation of calcite veins in releasing sections. In lenses between shear zones, homogeneous gouge is
formed with a well-developed oblique foliation and removal of calcite grains by pressure solution. We
infer that with progressive deformation, the number and density of slickensided shear surfaces increases,
generating tectonically derived scaly clay and more homogeneous gouge. In all deformed elements of the
Main Fault, porosity is much smaller than in the undeformed Opalinus Clay. An interesting observation is
the almost complete absence of cataclastic microstructures. Transmission electron microscopy (TEM) of
focused ion beam lamellae of this micron-wide shear zone shows a strong preferred orientation of clay
minerals, including nano-sized illite particles. In TEM, the shear zones envelop hard particles and confirm
an almost complete loss of porosity compared to the protolith.
We propose that inter- and transgranular microcracking, pressure solution, clay neoformation, crystal
plasticity and grain boundary sliding are important micro-scale processes during the early stages of
faulting in Opalinus Clay and thus need to be considered in extrapolating laboratory results to long-term
mechanical behavior.
© 2014 Elsevier Ltd. All rights reserved.
1. Introduction
http://dx.doi.org/10.1016/j.jsg.2014.07.014
0191-8141/© 2014 Elsevier Ltd. All rights reserved.
108 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
(e.g. Rutter et al., 1986; Logan et al., 1992; Haines et al., 2013). While undeformed fabric is heterogeneous and weakly foliated
the geometries of these shear zones can be similar, the micro-scale along bedding, with a porosity of 8e24% (Houben et al., 2013;
mechanisms of deformation can vary widely (e.g. Rutter et al., Nussbaum and Bossart, 2008).
1986), and the interplay of crystal plasticity in clays (Urai and (2) The detailed fracture network and strain intensity within
Wong, 1994), particle reorientation and pore collapse the zone of the Main Fault is heterogeneous (Nussbaum et al.,
(Morgenstern and Tchalenko, 1967; Milliken and Reed, 2010; 2011), comprising zones with fault gouge, C’-type shear
Haines et al., 2013), grain-size reduction (Bos and Spiers, 2001; bands (sensu Passchier and Trouw, 2005), meso-scale folds,
Mitra and Ismat, 2001; Rutter et al., 1986), solution-precipitation microfolds, numerous fault planes and apparently undis-
and grain-boundary sliding and mineral transformation (Boullier turbed parts (Fig. 2). Parts of the Main Fault (e.g. the upper
et al., 2009; Buatier et al., 2012a; Gratier et al., 2011; Haines and part of the outcrop in Fig. 2) comprise a ‘scaly’ fabric, where
van der Pluijm, 2012; Sasseville et al., 2012; Schleicher et al., the rock splits progressively into smaller fish-like flakes
2006; Warr and Cox, 2001) are all subjects of debate. (sensu Vannucchi et al., 2003).
Samples broken along these thin shear zones usually show (3) P-T conditions at the time of the onset of motion of the Main
slickensides, with a smooth, shiny striated surface (sensu Passchier Fault (late Miocene, Nussbaum et al., 2011) are inferred by
and Trouw, 2005). Slickensides are common in both experimentally Mazurek et al. (2006) to be about 55 C under an overburden
and naturally deformed samples of many materials and have been of 1000 m, using an integration of apatite fission track, vit-
studied by Doblas (1998), Gay (1970), Means (1987), Petit and rinite reflectance and biomarker isomerization analysis. The
Laville (1987) and Tjia (1964) and in clays by Dehandschutter maximum overburden and temperature of OPA at MT URL
et al. (2005), Gray and Nickelsen (1989), Labaume et al. (1997), was 1350 m and 85 C, respectively, during Cretaceous times
Saffer et al. (2012) and Will and Wilson (1989). (Mazurek et al., 2006). Thus, OPA, which now has a strongly
While the extreme localization of strain in a hierarchy of shear anisotropic unconfined compressive strength between 6 and
zones is generally recognized, the use of the terms “slip surface”, 28 MPa (Amann et al., 2011; Bock, 2001), was over-
“shear plane” and “thin slip zone” tends to differ between authors, consolidated at the onset of faulting.
and is scale dependent (e.g. Nussbaum et al., 2011; De Paola, 2013; (4) The kinematics of the Main Fault have been inferred by
Fondriest et al., 2013). Study of the grain-scale deformation paleostress analysis of slickenlines (Fig. 2), showing s1 to be
mechanisms in these zones is complicated by the very fine grain sub-horizontal, trending NNW-SSE for a reverse faulting
size and fragile fabric (cf. Vrolijk and van der Pluijm, 1999), which mode (Nussbaum et al., 2011). This geometry agrees with the
makes sample preparation and analysis difficult. In recent years, “distant push” or “Fernschub” theory of the Jura folding as a
ion- milling techniques have led to a breakthrough in sample consequence of the propagation of the Alpine foreland to-
preparation of clays, and to renewed interest in microstructural wards NNW (Laubscher, 1961). The Main Fault can be inter-
study (e.g. Holzer et al., 2006; Desbois et al., 2009, 2010, 2011). preted as a shear fault-bend fold (Nussbaum et al., 2011),
While the porosity and microfabric of tectonically undeformed which was passively steepened from 20 to the range of
Opalinus Clay (OPA) have been intensively studied (Houben et al., 40 e45 in sequence with the folding of the Mont Terri
2013; Keller et al., 2011; Wenk et al., 2008), little is known of the anticline over a basal ramp (Fig. 1). From area balancing
microstructure and deformation mechanisms in naturally and (Freivogel and Huggenberger, 2003), the offset of the Main
experimentally deformed OPA. Fault is inferred to be relatively small (~10 m).
This contribution reports an investigation of microstructures in (5) The paleo-fluid flux within the MT URL has been inferred
naturally deformed OPA from the Main Fault at the Mont Terri from calcite (CaCO3) and celestite (SrSO4) veins (Pearson
Underground Research Laboratory (MT URL, Fig. 1). We present et al., 2003). In a recent, comprehensive study, de Haller
microstructural data at scales ranging from dm to nm, obtained by et al. (2014) presented detailed petrographic and geochem-
light and electron microscopy. Based on these data, we discuss ical data about veins in the Main Fault. Veins were found as
deformation mechanisms in our samples, focusing on incipient thin (<1 mm thick) wafers of calcite and celestite, with
faulting, the early evolution of fault gouge, and processes for fibrous crack- seal microstructure, indicating syntectonic
resealing fluid pathways. precipitation. Based on isotope data (Sr, S, O, C), de Haller
et al. (2014) suggested that OPA acted as a seal for fluid
1.1. Geology flow during most of its history except during the movement
of the Main Fault. The present permeability of OPA is very low
As its main objective, the MT URL evaluates the long-term safety with no significant hydrological contrast between protolith
of radioactive waste disposal by basic and applied research. A and the Main Fault (~2 1013 m/s, (Nussbaum and Bossart,
detailed description on the local geological setting can be found in 2008). Profiles of a range of pore-fluid geochemical tracers
Nussbaum et al. (2011), in Bossart and Thury (2008), and in Becker are not perturbed near or within the Main Fault (Mazurek
(2000). The Main Fault, a 0.8e3 m wide fault zone exposed in the et al., 2011).
MT URL with a reverse offset of about 10 m, provides an excellent
opportunity to investigate incipient faulting in mudrocks, because 2. Samples and methods
of the unique access to extremely well preserved samples and the
large dataset of geochemical, structural and hydrological studies Drill cores from five boreholes in the fault zone (BSF-06, BPS-10,
(e.g. Pearson et al., 2003; Bossart and Wermeille, 2003; Nussbaum BPS-11, BPS-12 and BIC-A1, see Fig. 1 for location), up to 50 m apart,
and Bossart, 2008). Of this extensive database, the following as- were collected. Considering the extremely fragile nature of fault
pects are particularly relevant for this study: zones in clays, these cores were collected by drilling into the walls
and floor of the MT URL, with air as lubricant and careful core
(1) The protolith (sensu Rutter et al., 2001) is the shaly facies of extraction. Dip and reverse faulting mode of faults and slickenlines
OPA, which consists mainly of clay minerals (16e40% illite are in agreement with outcrop investigations by Nussbaum et al.
and 5e20% mixed-layer illite-smectite, 15e33% kaolinite, (2011). Strike orientations were not verifiable, as the drill cores
4e20% chlorite), 5e28% calcite (fossils and minor veins) and were not oriented. The cores BSF-06, BPS-10 and BPS-11 were taken
6e24% quartz (Pearson et al., 2003). The tectonically by a resin stabilized coring procedure (details in Nussbaum et al.,
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 109
Fig. 1. Geology of the Mont Terri URL. A) Facies map, borehole locations indicated. B) 2D balanced cross-section after Freivogel and Huggenberger (2003).
2006). Despite the care during drilling, the non-stabilized cores and parallel to the slickenlines, guided by a pre-cut, (ii) hand-
BPS-12 and BIC-A1 broke into pieces of up to 30 cm in length, along polishing of large (10 10 cm) surfaces followed by immersion in
shiny, wavy surfaces with striations (i.e. slickensides, Fig. 3). water for one second and drying to decorate foliation and shear
Although undeformed OPA produces quite strong and intact drill zones, (iii) ultra-thin sectioning (prepared by Geoprep, Basel), (iv)
cores, the collected core fragments easily broke along fractures hand polishing followed by broad ion beam (BIB) polishing, and (v)
with slickensides. cutting transmission electron microscopy (TEM) lamellae by
Preparation (dry) for this study used sample blocks of a few cm focused ion beam (FIB) (Fig. 2C).
on each side, cut from the core, which was resin-stabilized if Argon BIB polished sections of about 2 mm2 were prepared
required, with a low speed diamond saw. The samples have either using a JEOL SM 09010 (8.5 h at 6 kV, 150e200 mA, 103e104 Pa).
(1) a slickenside exposed on one side or (2) a slickenside inside the This process removed a layer about 100 mm thick from the pre-
block with the slickenside parallel to one side (Fig. 2C). The samples polished surface using the methods described in detail by
were investigated by optical and electron microscopy in surface Houben et al. (2013). The BIB polished, damage-free and flat
view (A) and in side view (B). In the latter case, a surface perpen- ( ± 5 nm, Klaver et al., 2012) surfaces allow detailed imaging of
dicular to the slickenside and parallel to the slickenlines was microfabric and pores. To prevent curtaining and edge effects, the
examined (Fig. 4). Here, the samples were further prepared by five slickensides were covered by gluing a 0.15 mm thick glass plate on
different methods: (i) breaking perpendicular to the slickenside the surface with epoxy resin before polishing (Fig. 4).
110
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 2. A) Paleostress analysis from striae (Nussbaum et al., 2011). B) Small-scale map of the Main Fault in gallery 08 (see Fig. 1 for location) modified after Nussbaum et al. (2011). C) Illustration of sample types and methods applied.
OM ¼ optical microscopy, SEM ¼ scanning electron microscopy, BIB ¼ broad ion beam milling, FIB ¼ focused ion beam milling, TEM ¼ transmission electron microscopy.
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 111
Fig. 3. Compilation of slickenside photographs from well BIC-A1. Sided light was used to minimize reflections. Sketches illustrate the surface orientation towards bedding foliation
(SO). Bold arrows indicate inferred movement of missing blocks. A) Sample BIC-A1-BL05 showing an intersection (branch line) of a wide stepped (1) and a narrow stepped (2)
slickenside. Arrows (a) and (b) point towards darker and brighter areas on the slickenside, respectively. B) Detail of sample BIC-A1-BL03, showing a tool track of ca. 1.6 cm, which
gives an estimate for the minimal offset along this slickenside. C) Detail of sample BIC-A1-BL03 illustrating polished vs. ragged parts. D) Slickenside from sample BIC-A1-BL06 with
large step length, but oblique to bedding.
Scanning electron microscopy (SEM) was performed on a Zeiss (using ArcGIS 10, ESRI 2013). The chosen cell size, being in the range
SUPRA 55 operating between 3 and 20 kV, using a backscattered of clay particles, gave the most practical spatial information.
electron detector (BSE) and a secondary electron detector (SE), For nm-scale investigation of the zone directly underlying the
respectively. Further, an energy-dispersive X-ray (EDX) detector (at slickenside, TEM was used to image FIB lamellae of 10 5 mm2
15e20 kV) was used. With these detectors, the depth in the sample and 100e150 nm thick (FIB: Strata 205, FEI, Central Facility for
from which information is provided is up to about 2 mm for the Electron Microscopy, University of Aachen, tungsten coating,
different detectors and acceleration voltages. 30 kV, 20 nAe100 pA; TEM: Zeiss Libra 200FE operating at
Following the procedures of Desbois et al. (2010), Houben et al. 200 kV, Koehler illumination system, Institute for Mineralogy,
(2013), Klaver et al. (2012) and Hemes et al. (2013), semi-automated University of Münster). High angle annular dark field (HAADF)
digital image analysis (DIA) was performed on high-resolution scanning TEM (STEM) and bright field (BF) images were collected,
mosaics of up to 400 single SE micrographs with magnifications giving information on average atomic number and mainly
up to 30 k. This approach provided segmented pores used for diffraction contrast, i.e. crystallinity of illuminated areas,
subsequent statistical analysis of size distribution, shape factors respectively. EDX measurements with short dwell times were
and orientation. Moreover, spatial distributions of these porosity used to minimize sample damage. Selected area electron
parameters were mapped. For each cell of a computed grid with a diffraction (SAED) patterns were recorded to specify major
cell size of 1 * 1 mm2, porosity, pore orientation and pore shape element characteristics and crystallographic parameters of re-
factors were calculated and each cell was colored accordingly gions of interest within the lamella.
112 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
3. Results
We use the term slickenside for the shiny, striated surface seen
on broken core pieces. The surface structure of slickensides shows
differences in step length, step height, shininess and orientation to
bedding. In the MT URL, the angle between the Main Fault zone and
bedding is about 15e25 and in our samples the angle between
slickensides and bedding foliation can vary between 0 and 45
(Fig. 3). Commonly, the step length is larger for slickensides parallel
to bedding, however exceptions occur (Fig. 3D).
At optical resolution, each slickenside consists of matt, ragged
areas and shiny, flat, polished areas, which show slight brightness
and color variations (Fig. 3A). By stereo microscopy, a few occur-
rences of asymmetric cavities, crescent- shaped markings and
trailed material were detected, all in agreement with reverse mo-
tion, such that the steps are congruous to shear sense (after the
terminology of Doblas, 1998). We found no offset markers to allow
measurement of displacement because this part of the Shaly Facies
is rather homogeneous without laminations that could be used as
markers. Based on a grain trace, the offset for one wide- stepped
(~3 cm) slickenside was determined to be at least 1.6 cm (Fig. 3B).
Curved fractures outcropping on a slickenside, sometimes with a
riser, were usually shown to be branch lines (sensu Walsh et al.,
1999) of intersecting slickensides (cf. Fig. 4). The vast majority of
striae resemble the continuous ridge-in-groove type described by
Means (1987), located next to risers, usually thinning out along the
steps (Fig. 3).
At SEM-scale, zoom-in sequences in Figs. 5 and 6 compare the
surfaces of narrow- to wide- stepped slickensides. The wide-
stepped slickenside has a smoother, less striated surface. For
scales of 100 mm and smaller, however, slickensides do not differ
much, and show at high magnifications particles of 200 nm and
less, surrounding larger particles (Figs. 5C,F and 6C). In BSE images
of a slickenside, brightness variations are interpreted as minerals
with a higher atomic number directly underlying clay particles
(Fig. 5E). Even at the limit of resolution of SEM (up to 100 k
magnification), porosity is not resolvable on these surfaces and the
slickensides appear as the surface of a continuous, smooth, non-
porous film covering underlying OPA. Occasionally, subeuhedral
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 113
Fig. 5. SEM images of a more narrow stepped (above) and a wider stepped (below) slickenside, samples BIC-A1-T1 and BIC-A1-U7, respectively. B) is an inset of A), in the lower left
the FIB-TEM location is visible. C) is an inset of B). E) is an inset of D) (small arrow pointing to location) showing a bright (heavy) mineral underlying a thin clay film in BSE mode,
20 kV. F) is an inset of E). All images (except E) are SE2 images; A)-C) 5 kV, D) 20 kV, F) 8 kV. Bold arrows indicate inferred movement of missing blocks. Note the clay particles with
sizes <200 nm at highest magnification.
calcite and celestite crystals are associated with risers (Fig. 6). In this image was taken on a surface which has been mechanically
BSE images, calcite and celestite lineations can be seen on some polished during preparation, and is not as smooth and damage-free
slickensides (Fig. 6). as a BIB-polished surface).
The few micron-wide zone is present immediately below all
3.2. Side view investigated slickensides. Fig. 8 shows SE images of a surface,
broken perpendicular to a slickenside in sample BIC-A1-U10. Here,
3.2.1. Thin zone underlying slickensides the undeformed OPA fabric is truncated by a sub-micron thick film
Fig. 7 shows a transmitted light micrograph of an ultra-thin showing a stacked sheet of overlapping platy clay particles that
section from sample BIC-A1-UTh6, with an open fracture that ex- wrap around a fossil grain.
poses slickensides on both fracture surfaces (shown in surface view Another illustration of this microstructure is shown in Fig. 9
in Fig. 3D). The fracture walls nest, indicating that no material fell (BIB-SEM sample BIC-A1-S9). Next to the slickenside surface, an
out from between them during later fracture opening. The image up to 3 mm thick zone is visible, with an internal microstructure of
was taken with crossed polarizers and gypsum plate, with the blue slickenside-parallel-oriented clay particles enveloping a pyrite
color (in the web version) showing the bedding-parallel preferred grain that slightly protrudes out of the slickenside. The amount of
orientation of clay particles of undeformed OPA. porosity of this sample is comparable to undeformed OPA (19 vs.
On both sides of Fig. 7A, a zone of higher order colors indicates 18e23%, cf. Houben et al., 2013) and lacks any trend in porosity,
enrichment of calcite in a 50 mm thick vein (Fig. 7C). On the left side pore shape or pore orientation up to the thin zone next to the
of the image, a fish-shaped zone is defined by sharp boundaries fracture (Fig. 9C and E). The thin zone, however, shows almost no
where the preferred orientation makes an angle with bedding porosity at SEM resolution (Fig. 9D and F). For imaging purposes,
foliation and seems to contain less of the bright calcite grains than the sample was prepared by gluing a glass plate on the slickenside
the surrounding protolith (Fig. 7B). At the interface between both before BIB polishing and epoxy could have filled pores in the thin
fracture walls, a few mm wide zone, barely visible by light micro- zone, so the next example shown is without epoxy.
scopy, differs from the surrounding material by color (in the web The absence of porosity in a thin zone is also visible in BIB-SEM
version) (Fig. 7D). The zone suggests a different preferred orienta- sample BIC-A1-U2 (Fig. 10). Here, an open, not epoxy filled fracture
tion than the protolith. In BSE imaging, this zone is darker, and (possibly revealing slickensides on both fracture walls) separates
more fine-grained, with a trace of an oblique foliation (note that the thin zone of aligned clay particles. Larger particles of calcite,
114 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 6. SEM images of sample BIC-A1-U5 in surface view (left) and side view (right). B) and C) are insets of A). Small arrows point to bright (celestite) lineations. Bold arrows indicate
movement of missing block. B) shows facetted celestite (Ce) next to clay minerals (Cl). C) shows nano-sized particles in high magnification (SE2, 5 kV). D) side view BSE image and E)
sketch, showing laminated calcite (Ca) and celestite (Ce) veins next to the slickenside. F) inset of D) picturing a BIB polished section. G) inset of F) showing clay minerals (Cl) in
between to laminated veins. See text for details. All images (except C) are BSE images, 20 kV.
quartz and pyrite framboids are never truncated by the fracture lattice spacing of 10Å (¼d001) (not shown). By STEM (HAADF),
walls. The thin zone is up to 2.5 mm thick, with almost no pores quartz grains close to the slickenside are found to be enveloped by
visible by SEM. A rare example of a tabular particle (mica) bent into bent illite grains (Fig. 12). The presence of nm-sized illite particles
parallelism with the slickenside is shown by the arrow (Fig. 10A). corresponds to the top-view observations presented above. Fig. 12B
The thin zone underlying the slickenside is examined in greater sketches the interpreted microstructure for this image, highlighting
detail in lamella TEM1 (Fig. 11), which was cut from the slickenside that the few small pores next to the slickenside are of inter-particle
shown in Fig. 5B. Therein, the ~1 mm thin zone consists of densely type and mainly related to pressure shadows of quartz grains. At
stacked clay particles, which are aligned (face to face texture) the same time, the finite thickness of the TEM lamellae causes the
parallel to the slickenside (Fig. 11B). By SAED and HRTEM, clay apparent porosity to be greater than seen on a 2D surface,
particles next to the slickenside were identified as illite, with a (Krabbendam et al., 2003).
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 115
Fig. 7. A) Ultra-thin section image (xpol, gypsum plate) of sample BIC-A1-UTh6. B), C) and D) are insets of A) picturing gouge, a vein and a thin band next to the fracture. E) is a BSE
micrograph (20 kV) of the sample location shown in D). F) is an inset of E) showing the bands thickness of approximately 4 mm. Qz ¼ quartz, Ca ¼ calcite, K-Fsp ¼ kalium feldspar,
Sid ¼ siderite.
In summary, FIB-TEM results agree with our BIB-SEM observa- the gouge contains less of the bright calcite particles than the
tions of a strong porosity reduction in a zone next to the slickenside protolith.
with a strong preferred orientation and the presence of nano- Sample BIC-A1-UThB4, similarly shows continuous gouge over
particles. Also, the grains in this zone and in this sample are illite, several cm, with a clearly lower content of bright calcite grains, a
suggesting that these zones can be locations of illite neoformation. sigmoidal foliation curving into parallelism with the lens boundary,
The transition from undeformed OPA into this zone is invariably and a darker color in parallel polarizers (Fig. 13B). The gouge
sharp, and occurs over a few 100 nm. thickness varies between 2 and 0.2 mm, being limited by one
almost straight and one irregular boundary to the protolith (Fig. 13).
3.2.2. Gouge A similar gouge lens is found in borehole BIC-A1 (Fig. 14). This
In Fig. 7A and B, we showed an example of the wider bands or sample was cut in two halves (BIC-A1-Th3 and BIC-A1-U8) with a
lenses of deformed material between some of the shear zones thin diamond saw, one half used for thin section and the other half
defining the slickensides. In this section, we describe these bands or (showing the same microstructure) for BIB-SEM analysis, locating
lenses in more detail. The ultra-thin section BIC-A1 UTh3 (xpol, the 1 mm2 of the BIB section to cover undeformed OPA, gouge and a
gypsum plate), has a few micron-wide zone branches enveloping a section of undisturbed contact (generally, gouge lenses easily
2 mm wide gouge lens (Fig. 13A). In this material, the preferred separate from protolith along slickensides). At the gouge e proto-
orientation makes an angle with the foliation of the protolith, and lith boundary, elongate particles are parallel to the boundary
116 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 8. Slickenside of broken sample BIC-A1-U10 in side view (SE2 image, 5 kV). Note the nano-sized width of the thin band next to the slickenside bordering OPA with clearly
undeformed bedding foliation (So). C is an inset of B (white rectangle).
(Fig. 14D), while in the gouge (Fig. 14C) there is a foliation in P- pores being more elongated than smaller ones. The orientations of
orientation (sensu Logan et al., 1979). At the boundary, a 1e2 mm the pores' longest axes are unimodally distributed and their mean
thin zone of nm-sized, boundary-parallel clay particles discor- orientation is in agreement with the minerals' P-orientation.
dantly covers the wall rock (Fig. 14D). The gouge shows a clear
decrease in grain size and an increase in fabric intensity compared 3.2.3. Veins and zones of calcite enrichment
to the protoliths' fabric (Figs. 14C and 9A). Particle sizes in the gouge In both optical microscopy and BIB-SEM, veins of calcite and
are less than 25 mm (usually less than 5 mm) in diameter, rounded sometimes celestite and zones of grain-scale calcite enrichment
and elongated and aligned along the gouge foliation (Fig. 14C). The were found, always located directly at the slickensides, both along
clay matrix shows a strong preferred orientation anastomosing the thin zones and the gouge-protolith contact (e.g. Figs. 7 and 15).
around more competent grains. Many particles are at the resolution Veins are thinner than 1 mm, commonly less than 0.1 mm thick. As
limit of BSE imaging. EDX mapping only resolves a change in shown above in surface view, calcite and celestite are occasionally
composition between protolith and gouge for a reduction in Ca due associated with risers and sometimes form slickenfibre veins (sensu
to the absence of large calcite grains (mainly fossils) in the gouge Passchier and Trouw, 2005) (Fig. 6A). This relationship is consistent
(e.g. Fig. 14B). Shear bands occur occasionally, defining an S-C fabric with the observations in side view: the BIB-SEM image of BIC-A1-
in the gouge (Fig. 14E). In a few cases, trans-granular cracks are U5 in Fig. 6FeG shows a thin clay layer separating two calcite-
visible in the larger grains, with crack porosity and clay filling in the celestite mixed veins. Such a vein texture is described by
crack (Fig. 14F). Passchier and Trouw (2005) as a striped shear vein with the clay
Segmentation of pores in SE images from the gouge shows a layer as an inclusion trail. The inclusion trail itself can be a slick-
porosity of less than 1% of inter-particle pores. This porosity is enside (cf. Stanley, 1990). From BIB-SEM, the porosity in veins is
much less than in undeformed OPA (8e24%, Houben et al., 2013). very low. The fine grain size of the veins prevents clear imaging in
The majority of pores within the gouge are elongated, with larger optical microscopy, but in BSE imaging (Fig. 15E and F), the veins are
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 117
Fig. 9. A) BSE micrograph (20 kV) of a resin (R) filled fracture in BIB-milled sample BIC-A1-S9. White arrows indicate nesting of a large pyrite grain (Py) into the opposite fracture
wall. B) inset of A) showing the alignment of clay particles next to the fracture and how they envelop the larger pyrite grain. Note the pores of visible in the SE micrographs C) and
D). E) and F) show the segmented pores (black) and the cell-wise porosity distribution (color intensity) for the same locations than A) and B). Qz ¼ quartz, Py ¼ pyrite. Note that
there is no visible trend in porosity towards the fracture wall, except from a thin zone (<3 mm) of almost no visible porosity next to the fracture. See text for details. (For inter-
pretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
shown to be laminated and, as the inset in Fig. 15A illustrates, terminating in the protolith (arrow in Fig. 15B). Veins usually have
contain grains elongated in the direction of the vein (cf. Koehn and at their boundary a thin shear zone (clay), which positions the veins
Passchier, 2000). Usually the fracture exposing the slickensides underneath the slickenside and which can occasional separate
runs along the contact protolith e vein on one side, but in some veins (Fig. 6C).
cases also internally, exposing a vein on both sides of the fracture The zones of grain-scale calcite enrichment are generally asso-
(Fig. 15A). Locally, the vein tip bends away from the slickenside, ciated with the veins, but can also be present at risers in the
118 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 10. BSE image (20 kV) of an open fracture separating a thin zone of aligned particles (sample BIC-A1-U2). B) is an inset of A). Arrow points to a tabular grain (probably mica)
bent into the thin zone of aligned clay minerals. Note that porosity is almost absent in the thin zone. Qz ¼ quartz, Ca ¼ calcite, M ¼ mica.
slickenside (Fig. 15C and D). These zones consist of a mixture of the sample during or after extraction from the drill core along zones
irregular- shaped calcite grains, sometimes showing zonation in with a very strong preferred orientation of fine grained clay. These
BSE, mixed with fine- grained clay fragments. zones are ubiquitous in samples from different locations of the
These observations are consistent with data from de Haller et al. Main Fault and usually not more than a few microns thick.
(2014), who studied vein fragments extracted from disintegrated Considering the very low tensile strength of OPA in a direction
samples of the Main Fault. We show that the veins are always perpendicular to the bedding-parallel foliation (Bock, 2001), it is
closely associated with slickensides, and form thin patches along not surprising that the cores fracture so easily along these zones.
these surfaces. We interpret the zones underneath the slickensides as thin
shear zones, along which strongly localized deformation produced
3.2.4. Scaly clay the observed fabric. With progressive deformation the shear zones
Parts of the outcrop shown in Fig. 2 and drill cores BIC-A1 and can accommodate limited displacement because of their curvature
BPS-12 show zones of ‘scaly’ fabric (Fig. 16). These zones range in before new shear zones nucleate (cf. Fig. 2 of Ingram et al., 1997). In
width from several cm up to ca. 1 m. There, the rock is separated the samples with widely-spaced shear zones, the surrounding OPA
into progressively smaller fish-like flakes down to less than 300 mm is only weakly deformed, as shown by the cusps in the fossils
in diameter, each of which is bound by slickensides. In this interpreted to represent minor pressure solution (Den Brok and
contribution, we present a first look at a small number of these Morel, 2001). In some other cases, lenses enclosed between shear
flakes, embedded in epoxy and polished by BIB. The samples zones are deformed more homogeneously (Van der Zee and Urai,
studied so far contain an internal fabric similar to the protolith, 2005) to produce gouge. With progressive deformation, the den-
intersected with incipient shear zones (Fig. 17). The rim of the flake sity of shear zones increases, forming an anastomosing network of
shows a 1e2 mm thin zone of parallel oriented clay particles, scaly clay. Our first results show that this network occurs, inter-
similarly found in thin zones underlying slickensides from the non- estingly, without much deformation inside the lenses. A perhaps
scaly clay part of the outcrop. More work is needed to test if this is surprising result of our observations is the absence of tips for the
representative for all scaly clay in the Main Fault, but these data slickensides, which indicates that they propagated long distances
provide a first impression that the scaly clays are defined by a for very small bulk deformations, terminating at branch lines in the
higher density of shear zones, and not by internal deformation of shear zone network.
the flakes resembling microlithons.
4.1. Thin shear zones
4. Discussion
Although we did not find indicators of absolute displacement on
The outcrops of the Main Fault in the MT URL provide a unique individual slickensides and direct indicators of shearing like grains
opportunity to study exceptionally well preserved samples bending into the shear zone are rare, we interpret the shear zones
recording the early stages of faulting in a slightly indurated clay- to have displacements in the mm- to cm range (e.g. inferred from
stone. Timing and conditions of the faulting are reasonably con- the tool track in Fig. 3B). The data seem to indicate a relation be-
strained and to date the fault zone is known to have a permeability tween the slickenside e bedding angle and step length. We hy-
as low as the surrounding protolith. Motion of the Main Fault could pothesize that a relationship between slickenside morphology and
be associated with small seismic slips, triggered in the hard Triassic displacements also exists, with surfaces like number (2) in Fig. 3
dolostones below or by stress changes during motion of the basal representing small displacements, before the individual segments
thrust, and followed by the much weaker OPA. coalesce into a fully mobilized shear zone.
Of the structural elements described in this paper, Fig. 18 sum- The internal microstructure of the thin shear zones (<4 mm
marizes the main features in a generic model. Slickensides are the width) is characterized by (i) very strong preferred particle orien-
main structural element in the samples studied. It is commonly tation, (ii) extremely low porosity, (iii) bending around larger
agreed that they were formed by localized shearing of the protolith grains, (iv) possibly more illite than in undeformed OPA (based on
in many lithologies and under a range of conditions. From our ob- one FIB-TEM sample), (v) the presence of nanoparticles, and (vi)
servations at the sub-micron scale slickensides are similar in almost no transition zone from the undeformed fabric to the shear
morphology in the Main Fault. They are revealed by fracturing of zone.
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 119
Fig. 11. A) STEM (HAADAF) mosaic of FIB lamella TEM1. B) is an inset of A), showing face-to-face aligned clay particles. C) is a sketch of B). D) is an inset of A), showing bent illite
particles enveloping quartz grains. E) sketch of D). I ¼ illite, M ¼ mica, Qz ¼ quartz, see text for details.
Tool tracks, such as the grain trace in Fig. 3B, indicate frictional contributing to the waviness of the shear zone. Thus, in surface
sliding in the shear zones (Hancock and Barkaa, 1987). Although we view the continuously connected, smooth clay film is seen (Fig. 5).
do not know the shear strength of these shear zones under con- Frictional sliding and associated abrasive wear are common
ditions of natural deformation, the residual shear friction angle of mechanisms for the generation of gouge and the mirror-like pol-
Opalinus Clay, which has been measured to be about 22 (Bock, ishing of slickensides in brittle environments (Blenkinsop, 2000;
2001), is a safe upper bound for this. Therefore, these shear Fondriest et al., 2013; Hancock and Barkaa, 1987; Tjia, 1964; Twiss
zones, once formed, are expected to localize deformation and to and Moores, 1992). In claystones, this process can produce soft
propagate rapidly (Collettini et al., 2009). clay gouge even from strongly cemented shales (Holland et al.,
The shear zones anastomose around larger minerals and do not 2006). These mechanisms are commonly associated with inter-
develop enough stress to fracture these grains (Jessell et al., 2009), and trans-granular microcracking (Niemeijer et al., 2010). However,
120 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 12. Detail STEM (HAADAF) image of D, showing nano-sized illite particles, some examples marked by arrows.
trans-granular microcracks are almost completely absent in our ductile deformation of the protolith, dissolution of grains and pore
shear zones, in agreement with results from other clay gouges collapse being important micro-scale processes (cf. Bos, 2002;
(Haines et al., 2013; Milliken and Reed, 2010). Niemeijer and Spiers, 2007, 2005; Niemeijer et al., 2008). This
This absence may indicate the importance of other mechanisms interpretation is also in agreement with the roughly similar
such as grain-boundary sliding, crystal plasticity of phyllosilicates, composition of protolith and gouge matrix around larger grains
solution-precipitation processes or neoformation of clay minerals, measured by SEM-EDX. Evidence for cracking of grains and clay
associated with microscopically ductile creep. Slip on the (001) filling in the fractures is present but rare (Fig. 14F).
basal planes of clay particles is much easier than shearing related to The 1e2 mm thin shear zones, containing nm-sized clay minerals
grain breakage (cf. Haines et al., 2013). The abundant nanoparticles are present at the gouge zone boundaries (Y-shears sensu Logan
in the shear zone could have been formed by both, cataclasis and et al., 1979). We suggest that the gouge forms by deformation of
solution-precipitation processes, by recrystallization of kinked and lenses between thin shear zones, driven by the stress increase in
folded clay mineral grains (Urai et al., 1980) and by neoformation of the lenses during deformation of the weak micro-shear zones (Van
illite (Vrolijk and van der Pluijm, 1999). Although the smectite- der Zee and Urai, 2005). This interpretation is in agreement with
eillite transformation has been shown to occur preferentially in the sigmoidal foliation in the gouge lenses.
fault gouge (Vrolijk and van der Pluijm, 1999), we do not have Still, the micromechanisms driving the gouge evolution are
sufficient evidence for this process, as we only analyzed one TEM difficult to identify. The removal of calcite inside the gouge and the
sample so far. However, the densely stacked illite aggregates in the high fabric intensity indicates pressure-solution in the lenses be-
shear zone of Fig. 11E suggest that smectite-illite transformation tween thin shear zones. We hypothesize that particle comminution
could have occurred but this hypothesis should be further tested by was a relevant micromechanism. However, even though individual
analysis of the minerals in the micro shear zone and in the protolith grains at the host rock boundaries seem truncated by the thin shear
(cf. Buatier et al., 2012b; Sasseville et al., 2012). zone (quartz grain on the left in Fig. 14D), we cannot identify their
The actual processes of localization of strain during propagation missing counterparts neither inside the gouge nor in the host rock
of the shear zones and their relative importance could be studied at at the other side of the gouge. The fact that trans-granular micro-
the tip of a propagating shear zone. Unfortunately, these features cracks inside the gouge are rarely observed, might be explained by a
were not found so far. The reorientation of clay grains, fluid gouge internal plastic deformation to maintain the P-orientation
expulsion during pore collapse, formation of preferred particle under ongoing deformation (cf. Haines et al., 2013). This process
orientation and strong grain-size reduction by cataclasis could all could lead to homogenization of the gouge microstructure, making
have been dominant mechanisms in the deformation of OPA. it difficult to relate broken grain parts. The strong P-orientation of
Testing this hypothesis requires further detailed analysis of grains in gouge (S-C fabric, Figs. 13 and 14) has been described in
mineralogy, microchemistry and isotope compositions. nature and experiments (e.g. Sibson, 1977; Yan, 2001; Zulauf et al.,
1990; Logan et al., 1979; Rutter et al., 1986; Haines et al., 2013). The
4.2. Gouge total offset of the gouge bands might lay in the range of several cm
to dm, with the thin shear zones at the gouge boundaries likely
The microstructure of gouge is very different from that of the accommodating the majority of this offset, acting as Y-shears (cf.
thin shear zones, defining the boundaries of the gouge. The gouges' Logan and Rauenzahn, 1987). The absence of intra-granular
strong foliation defined by clay minerals anastomosing around microfractures and the strong particle alignment within the thin
larger, rounded, elongated grains of quartz and calcite, the very low shear zones indicates stable sliding. This proposed process is in
porosity, the absence of large calcite grains all agree with strong, agreement with the interpretation of Y-shears as planes of stable
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 13. Transmitted light microscope images of ultra-thin sections BIC-A1-UTh3 (A) and BIC-A1-UThB4 (B) (xpol, gypsum plate), showing lenses of gouge with sigmoidal particle orientations and darker color compared to surrounding
protolith. Arrows give inferred sense of movement. See text for details. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
121
122 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 14. A) is an optical micrograph of thin section BIC-A1-Th3 showing a gouge lense (darkbrown). B) inset of A) showing a calcite fossil with a straight edge at the gouge e protolith
boundary. C) shows gouge internal fabric. D) pictures the sharp gouge e protolith boundary (between bold arrows). E) shows a thin shear zone in the gouge. F) shows a cracked
quartz grain in the gouge. C), D), E) and F) (all BSE images, 20 kV) derive from sample BIC-A1-U8. Comparable locations are indicated in A), with E) and F) close to C), see text for
details. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
sliding (Logan and Rauenzahn, 1987), which form subsequent to shearing (Koehn and Passchier, 2000), with growth of calcite into
work hardening (e.g. accompanied by stress increase in lenses as dilation sites. Dilatancy is consistent with a slight over-
stated above) and fabric change that, in our case, could be attrib- consolidation of OPA (Cuss et al., 2011; Ingram and Urai, 1999)
uted to an initial comminution process. We cannot judge whether during the evolution of the Main Fault (Nussbaum et al., 2011).
this feature represents a geological marker for seismic slip, as Klinkenberg et al. (2009) reported the occurrence of inter-particle
stated by Fondriest et al. (2013) for shears in dolostone gouge, or for cracks connecting larger grains in the surrounding of shear zones
sub-critical (but repeated) crack-grow. The latter would match with in synthetically deformed OPA. Based on unconfined uniaxial
a contemporaneous pressure solution process of Ca (section above). compression tests on OPA, Amann et al. (2011) proposed the for-
Moreover, a mechanical comminution could result in a decreased mation of microcracks around larger heterogeneities as a reason for
dissolution contact area, which, in turn, enhances the pressure failure far before rupture stress.
solution process of Ca described above (cf. Gratier et al., 2014). Crack-seal vein textures are commonly interpreted to indicate
incremental opening of the crack, with crack apertures in the
4.3. Veins micron range. Calcite and celestite veins, clearly visible in thin
section (Fig. 15A and B), are interpreted to cause the brightness
Veins are thin, locally fibrous and laminated (Fig. 15E and F), variances on the slickensides (Fig. 3A). Laterally, they pass into the
which is a texture proposed to indicate crack-seal processes during shear zone without veins, suggesting a distribution of veins as
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 123
Fig. 15. A) and B) are optical micrographs of ultra-thin section BIC-A1-UTh6 (xpol, gypsum plate) showing sharp vein e fracture boundaries with veins on both fracture sides in A)
and a vein just on one fracture side in B). Arrow is pointing to a vein tip terminating in the protolith. C) is a reflected light micrograph from sample BIC-A1-U7 (sided light)
illustrating calcite enrichment at a riser. D) is an inset of C). E) shows laminated veins (Lam.), for location see Fig. 7A. F) pictures a laminated vein at the gouge-protolith boundary,
see Fig. 13B for location. D), E) and F) are BSE images taken at 20 kV. Ca ¼ calcite, Qz ¼ quartz.
patches in the shear zone. The local, occasional presence of euhe- curvature can be expected to lead to dilatancy and drop in fluid
dral, faceted grains indicates growth into free fluid and the pres- pressure in releasing sections and simultaneous precipitation of
ence of local open cracks. calcite. This process indicates a mixed-mode failure generating
Because of the slight curvature of the shear zones, shearing in- hybrid extensional-shear fractures (sensu Sibson, 2000). The calcite
duces stress changes between releasing and restraining sections was perhaps generated by dissolution in the gouge and migration
(Bürgmann and Pollard, 1992; Chester and Chester, 2000; Van der through local cracks in the otherwise non-porous thin shear zones
Zee and Urai, 2005). In the slightly overconsolidated OPA, this bordering the gouge lenses or came from an external source. Veins
124 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Fig. 16. A) is a photograph (sided light) of water immersed hand specimen BPS12-03, showing a distinct boundary of bedding foliation to scaly clay fabric along a sharp shear zone
(dashed line in sketch B). C) and D) are photo and sketch of an inset in A).
may locally restore the strength of some shear zones (Wintsch, disrupted veins (cf. Vrolijk and Sheppard, 1991). Fig. 15E and F
1998), but also provide regions of stress localization and propaga- shows a somewhat disordered vein structure supporting a tectonic
tion of shear. On the other hand, veins may also reflect higher fluid reassembling. Whether this explanation applies for calcite enrich-
pressures indicating dilatant fracturing. ments in a releasing section such as the riser in Fig. 15C is unclear.
In our interpretation, the veins formed in evolving shear zones,
contemporaneous with the reverse shearing, but after an initial 4.4. Scaly clay
nucleation of the shear zones. This interpretation is supported by
the laminated microstructure (Fig. 15) coherent with reverse Tectonically derived scaly clay (Cowan et al., 1984; Labaume
shearing of the shear zones. et al., 1997; Lundberg and Moore, 1986; Maltman, 1998; Moore
The zones of grain-scale calcite enrichment (Fig. 15C), also and Byrne, 1987; Vannucchi et al., 2003) in the Main Fault de-
interpreted to have formed in releasing sections and associated velops in C’-type shear bands, but also in meso-scale folds and
with veins, are less easily explained. Our preferred explanation is microfolds (Nussbaum et al., 2011). As suggested by the first results
that these are small relay zones, which failed in extension during of this study, scaly clay consists of small, fish-shaped flakes of un-
early linkage of segments of the forming shear zones by distributed deformed or weakly deformed OPA bound by slickensides. With the
microcracking and simultaneous precipitation of calcite in the increase in density of shear zones, an anastomosing network de-
microcracks. In agreement with this, Vrolijk and Sheppard (1991, velops, leading to macroscopically ductile shearing. An interesting
Fig. 11) show isolated fibrous carbonate spheroids, possibly indi- question here is the dissipation of excess pore pressure that is
cating incipient filling of a fracture. Another explanation is that locally created by collapse of the porosity during shear zone
these zones of grain-scale calcite enrichment are tectonically propagation: with widely spaced shear zones this dissipation is
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 125
Fig. 17. A) is a reflected light microscope image of scaly clay flakes showing slickensides. B) is a high magnification SE (5 kV) micrograph showing nano-sized particles on a
slickenside from a scaly clay flake (sample BPS12-3b). C) BSE micrograph of BIB-milled scaly clay flake suspended in resin. D) detail of C) showing a thin, non-porous zone of aligned
particles at the flakes boundary.
Fig. 18. Schematic model of faulted OPA from the Main Fault in the MT URL. Bold arrows indicate movement of the missing block.
126 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
much easier than in a volume bound by a network of non-porous Calcite and celestite veins occur along some shear zones. Tex-
shear zones that presumably form permeability barriers, perhaps tures of the veins indicate crack-seal growth and only locally
providing a feedback mechanism to generate local high density of growth into a free fluid. Calcite enrichments are often associated
shear zones and scaly clay. This interpretation is supported by Arch to risers of the slickensides.
et al. (1988), who demonstrated that (1) small differences in water Elongated and lenses of gouge (<1 cm thick) are also found in
content can produce a large difference in both; the peak and re- the Main Fault, bound between thin shear zones. Gouge is
sidual strengths and that (2) higher water content leads to strongly but more homogeneously deformed, with grain size
increased complexity in shear zone geometry in clays. and porosity reduction, removal of calcite grains by pressure
solution and development of a tectonic foliation. We interpret
4.5. Implications to present-day permeability and paleo-fluid flow the gouge to have formed in restraining bends between shear
of the Main Fault zones, and the veins in releasing bends.
Microstructural evidence for cataclasis is rare.
Tectonically undisturbed Opalinus clay has a very low perme-
ability (2 1013 m/s, Bossart and Wermeille, 2003). One of the Acknowledgments
most important results of this study for fluid flow and self- sealing
models (e.g. Bock et al., 2005) is that in-situ the Main Fault contains The authors like to thank Norbert Clauer, Isabelle Techer and
a network of shear zones, veins and gouge with a much lower Sam Haines for valuable discussions. Willy Tschudin is thanked for
porosity than the protolith. This outcome is consistent with the his extraordinary “art” of ultra-thin section preparation. We thank
measurement of no permeability increase in the Main Fault Editor Bill Dunne, Peter Vrolijk and an anonymous reviewer for
(Nussbaum and Bossart, 2008), but predicts that the present-day their constructive feedback. This work is funded by the Mont Terri
permeability of the Main Fault is much lower than that of the Consortium of Swisstopo (CH).
protolith. However, this present-day permeability can also be easily
increased by microcracking along the very weak shear zones, e.g.
caused by excavation, tectonic forces, or fluid pressure. References
On the other hand, the occurrence of veins in shear zones are
indictors for paleofluid flow along releasing segments of the shear Agar, S.M., Prior, D.J., Behrmann, J.H., 1989. Back-scattered electron imagery of the
tectonic fabrics of some fine-grained sediments: implications for fabric
zones, which localized fluid flow and finally contributed to nomenclature and deformation processes. Geology 17, 901e904.
resealing the fault by precipitation of calcite and celestite. Based on Amann, F., Button, E.A., Evans, K.F., Gischig, V.S., Blümel, M., 2011. Experimental
isotope profiles (Sr, S, O, C), de Haller et al. (2014) suggested that study of the brittle behavior of clay shale in rapid unconfined compression.
Rock Mech. Rock Eng. 44, 415e430. http://dx.doi.org/10.1007/s00603-011-
OPA acted as a seal for fluid flow during most of its history except 0156-3.
during the movement of the Main Fault. This interpretation is in full Arch, J., Maltman, A.J., Knipe, R.J., 1988. Shear-zone geometries in experimentally
agreement with our results. deformed clays: the influence of water content, strain rate and primary fabric.
J. Struct. Geol. 10, 91e99.
Becker, A., 2000. The Jura Mountains e an active foreland fold-and-thrust belt?
4.6. Mechanical properties of the Main Fault during tectonic Tectonophysics 321, 381e406.
Blenkinsop, T.G., 2000. Deformation Microstructures and Mechanisms in Minerals
deformation and Rocks. Kluwer Academic Publishers, Dordrecht; Boston.
Bock, H., 2001. Rock Mechanics Analyses and Synthesis (RA Experiment): Data
The inferred fluid-assisted pressure solution and neo- Report on Rock Mechanics. Mont Terri Technical Report 2000e02. Bern-Ittigen.
Bock, H., Blümling, P., Konietzky, H., 2005. Study of the micro-mechanical behaviour
crystallization processes during tectonic deformation of the Main
of the Opalinus clay: an example of co-operation across the ground engineering
Fault are unlikely to occur in rapid laboratory experiments, but can disciplines. Bull. Eng. Geol. Environ. 65, 195e207. http://dx.doi.org/10.1007/
be simulated by analog experiments (Bos and Spiers, 2001; cf. s10064-005-0019-9.
Niemeijer et al., 2008). Bos, B., 2002. Frictional-viscous flow of phyllosilicate-bearing fault rock: micro-
physical model and implications for crustal strength profiles. J. Geophys. Res.
This matter means that the Main Fault during tectonic defor- 107, 2028. http://dx.doi.org/10.1029/2001JB000301.
mation had a different constitutive behavior than measured in the Bos, B., Spiers, C., 2001. Experimental investigation into the microstructural and
laboratory, and extrapolation of laboratory-derived mechanical mechanical evolution of phyllosilicate-bearing fault rock under conditions
favouring pressure solution. J. Struct. Geol. 23, 1187e1202. http://dx.doi.org/
properties to predict long-term creep deformation could be incor- 10.1016/S0191-8141(00)00184-X.
rect. The extrapolation of laboratory-derived mechanical properties Bossart, P., Thury, M. (Eds.), 2008. Mont Terri Rock Laboratory: Project, Programme
is a good upper bound and during long-term creep faults in OPA are 1996 to 2007 and Results. Wabern.
Bossart, P., Wermeille, S., 2003. Paleohydrological study on the surroundings of the
predicted to be weaker and more viscous. Mont Terri rock laboratory. In: Heitzmann, P., Tripet, J.-P. (Eds.), Mont Terri
Project e Geology, Paleohydrology and Stress Field of the Mont Terri Region,
Geology Series. Federal Office for Water and Geology, Bern-Ittigen, pp. 45e64.
5. Conclusions Boullier, A.-M., Yeh, E.-C., Boutareaud, S., Song, S.-R., Tsai, C.-H., 2009. Microscale
anatomy of the 1999 Chi-Chi earthquake fault zone. Geochem. Geophys. Geo-
Thin, localized shear zones (<4 mm width) are the elementary syst. 10 http://dx.doi.org/10.1029/2008GC002252.
Buatier, M.D., Chauvet, A., Kanitpanyacharoen, W., Wenk, H.R., Ritz, J.F., Jolivet, M.,
building blocks of the Main Fault, as they form an inter- 2012a. Origin and behavior of clay minerals in the Bogd fault gouge, Mongolia.
connected network through otherwise undeformed host rock. J. Struct. Geol. 34, 77e90. http://dx.doi.org/10.1016/j.jsg.2011.10.006.
The shear zones comprise aligned and enveloped particles as Buatier, M.D., Lacroix, B., Labaume, P., Moutarlier, V., Charpentier, D., Sizun, J.P.,
Trave , A., 2012b. Microtextural investigation (SEM and TEM study) of phyllo-
well as nano-sized illites and have a dramatically reduced silicates in a major thrust fault zone (Monte Perdido, southern Pyrenees):
porosity compared to the protolith. impact on fault reactivation. Swiss J. Geosci. 105, 313e324. http://dx.doi.org/
Samples fracture preferentially along these shear zones, 10.1007/s00015-012-0098-0.
Bürgmann, R., Pollard, D.D., 1992. Influence of the state of stress on the brittle-
revealing highly polished slickensided surfaces. These fractures
ductile transition in granitic rock: evidence from fault steps in the Sierra
were however not open in the samples before coring. Nevada, California. Geology 20, 645. http://dx.doi.org/10.1130/0091-7613(1992)
Tectonically derived scaly clay comprises fish-shaped flakes of 020<0645:IOTSOS>2.3.CO;2.
undeformed OPA of less than 300 mm thickness bound by thin Chester, F.M., Chester, J.S., 2000. Stress and deformation along wavy frictional faults.
J. Geophys. Res. 105, 23421. http://dx.doi.org/10.1029/2000JB900241.
shear zones. Scaly clay developed by progressive increase of the Collettini, C., Niemeijer, A., Viti, C., Marone, C., 2009. Fault zone fabric and fault
density of thin shear zones. weakness. Nature 462, 907e910. http://dx.doi.org/10.1038/nature08585.
B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128 127
Cowan, D., Moore, J., Roeske, S.M., Lundberg, N., Lucas, S.E., 1984. Structural features Jessell, M.W., Bons, P.D., Griera, A., Evans, L. a., Wilson, C.J.L., 2009. A tale of two vis-
at the deformation front of the Barbados Ridge complex, Deep Sea Drilling cosities. J. Struct. Geol. 31, 719e736. http://dx.doi.org/10.1016/j.jsg.2009.04.010.
Project LEG 78A. Initial Rep. Deep Sea Drill. Proj. 78, 535e548. http://dx.doi.org/ Keller, L.M., Holzer, L., Wepf, R., Gasser, P., 2011. 3D geometry and topology of pore
10.2973/dsdp.proc.78a.127.1984. pathways in Opalinus clay: implications for mass transport. Appl. Clay Sci. 52,
Cuss, R.J., Milodowski, A., Harrington, J.F., 2011. Fracture transmissivity as a function 85e95. http://dx.doi.org/10.1016/j.clay.2011.02.003.
of normal and shear stress: first results in Opalinus clay. Phys. Chem. Earth Parts Klaver, J., Desbois, G., Urai, J.L., Littke, R., 2012. BIB-SEM study of the pore space
A/B/C 36, 1960e1971. http://dx.doi.org/10.1016/j.pce.2011.07.080. morphology in early mature Posidonia shale from the Hils area, Germany. Int. J.
de Haller, A., Mazurek, M., Spangenberg, J., Mo €ri, A., 2014. Self-sealing of Faults (SF) Coal Geol. 103, 12e25. http://dx.doi.org/10.1016/j.coal.2012.06.012.
Project : Final Report, Mont Terri Technical Report. Federal Office of Water and Klinkenberg, M., Kaufhold, S., Dohrmann, R., Siegesmund, S., 2009. Influence of
Geology, Bern-Ittigen. carbonate microfabrics on the failure strength of claystones. Eng. Geol. 107,
De Paola, N., 2013. Nano-powder coating can make fault surfaces smooth and shiny: 42e54. http://dx.doi.org/10.1016/j.enggeo.2009.04.001.
implications for fault mechanics? Geology 41, 719e720. http://dx.doi.org/ Koehn, D., Passchier, C.W., 2000. Shear sense indicators in striped bedding-veins.
10.1130/G31398.1.Di. J. Struct. Geol. 22, 1141e1151. http://dx.doi.org/10.1016/S0191-8141(00)00028-6.
Dehandschutter, B., Vandycke, S., Sintubin, M., Vandenberghe, N., Wouters, L., 2005. Krabbendam, M., Urai, J.L., van Vliet, L.J., 2003. Grain size stabilisation by dispersed
Brittle fractures and ductile shear bands in argillaceous sediments: inferences graphite in a high-grade quartz mylonite: an example from Naxos (Greece).
from Oligocene Boom Clay (Belgium). J. Struct. Geol. 27, 1095e1112. http:// J. Struct. Geol. 25, 855e866. http://dx.doi.org/10.1016/S0191-8141(02)00086-X.
dx.doi.org/10.1016/j.jsg.2004.08.014. Labaume, P., Maltman, A.J., Bolton, A., Tessier, D., Ogawa, Y., Takizawas, S., 1997. Scaly
Den Brok, S.W.J., Morel, J., 2001. The effect of elastic strain on the microstructure of fabrics in sheard clays from the decollement zone of the Barbados accretionary
free surfaces of stressed minerals in contact with an aqueous solution. Geophys. prism. In: Shipley, T., Ogawa, Y., Blum, P., Bahr, J. (Eds.), Proceedings of the Ocean
Res. Lett. 28, 603e606. http://dx.doi.org/10.1029/2000GL011461. Drilling Program, pp. 59e77.
Desbois, G., Urai, J.L., Kukla, P. a, 2009. Morphology of the pore space in claystones e Laubscher, H., 1961. Die Fernschubhypothese der Jurafaltung. Eclogae Geol. Helv. 54,
evidence from BIB/FIB ion beam sectioning and cryo-SEM observations. eEarth 221e280.
4, 15e22. http://dx.doi.org/10.5194/ee-4-15-2009. Logan, J.M., Rauenzahn, K.A., 1987. Frictional dependence of gouge mixtures of
Desbois, G., Urai, J.L., De Craen, M., 2010. In-situ and Direct Characterization of quartz and montmorillonite on velocity, composition and fabric. Tectonophysics
Porosity in Boom Clay ( Mol Site, Belgium ) by Using Novel Combination of Ion 144, 87e108.
Beam Cross-sectioning, SEM and Cryogenic Methods. External Report of the Logan, J.M., Friedman, M., Higgs, N., Dengo, C., Shimamoto, T., 1979. Experimental
Belgian Nuclear Research Centre. Mol. studies of simulated gouge and their application to studies of natural fault
Desbois, G., Urai, J.L., Kukla, P. a., Konstanty, J., Baerle, C., 2011. High-resolution 3D zones. In: Proceedings of Conference VIII on Analysis of Actual Fault Zones in
fabric and porosity model in a tight gas sandstone reservoir: a new approach to Bedrock. US Geological Survey, Open File Report, pp. 79e1239.
investigate microstructures from mm- to nm-scale combining argon beam Logan, J.M., Dengo, C.A., Higgs, N.G., Wang, Z.Z., 1992. Fabrics of experimental fault
cross-sectioning and SEM imaging. J. Pet. Sci. Eng. 78, 243e257. http:// zones: their development and relationship to mechanical behavior. In:
dx.doi.org/10.1016/j.petrol.2011.06.004. Evans, B., Wong, T. (Eds.), Fault Mechanics and Transport Properties of Rocks e a
Doblas, M., 1998. Slickenside kinematic indicators. Tectonophysics 295, 187e197. Festschrift in Honor of W. F. Brace. Academic Press, pp. 33e67. http://dx.doi.org/
http://dx.doi.org/10.1016/S0040-1951(98)00120-6. 10.1016/S0074-6142(08)62814-4.
Fondriest, M., Smith, S. a. F., Candela, T., Nielsen, S.B., Mair, K., Di Toro, G., 2013. Lundberg, N., Moore, C., 1986. Macroscopic structural features in Deep Sea drilling
Mirror-like faults and power dissipation during earthquakes. Geology 41, project cores from forearc regions. Geol. Soc. Am. Mem. 166, 13e44.
1175e1178. http://dx.doi.org/10.1130/G34641.1. Maltman, a. J., 1998. Deformation structures from the toes of active accretionary
Freivogel, M., Huggenberger, P., 2003. Modellierung bilanzierter Profile im Gebiet prisms. J. Geol. Soc. Lond. 155, 639e650. http://dx.doi.org/10.1144/
Mont Terri e La Croix (Kanton Jura). In: Heitzmann, P., Tripet, J.-P. (Eds.), Mont gsjgs.155.4.0639.
Terri Project e Geology, Paleohydrology and Stress Field of the Mont Terri Region, Mazurek, M., Hurford, A.J., Leu, W., 2006. Unravelling the multi-stage burial history
Geology Series. Federal Office for Water and Geology, Bern-Ittigen, pp. 7e44. of the Swiss Molasse Basin: integration of apatite fission track, vitrinite
Gay, N., 1970. The formation of step structures on slickensided shear surfaces. reflectance and biomarker isomerisation analysis. Basin Res. 18, 27e50. http://
J. Geol. 78, 523e532. dx.doi.org/10.1111/j.1365-2117.2006.00286.x.
Gratier, J.-P., Richard, J., Renard, F., Mittempergher, S., Doan, M.-L., Di Toro, G., Mazurek, M., Alt-Epping, P., Bath, A., Gimmi, T., Waber, H.N., Buschaert, S., De
Hadizadeh, J., Boullier, a.-M., 2011. Aseismic sliding of active faults by pressure Cannie re, P., De Craen, M., Gautschi, A., Savoye, S., Vinsot, A., Wemaere, I.,
solution creep: evidence from the San Andreas fault observatory at depth. Wouters, L., 2011. Natural tracer profiles across argillaceous formations. Appl.
Geology 39, 1131e1134. http://dx.doi.org/10.1130/G32073.1. Geochem. 26, 1035e1064. http://dx.doi.org/10.1016/j.apgeochem.2011.03.124.
Gratier, J.-P., Renard, F., Vial, B., 2014. Postseismic pressure solution creep: evidence Means, W.D., 1987. A newly recognized type of slickenside striation. J. Struct. Geol.
and time-dependent change from dynamic indenting experiments. J. Geophys. 9, 585e590.
Res. Solid Earth 119, 2764e2779. http://dx.doi.org/10.1002/2013JB010768. Milliken, K.L., Reed, R.M., 2010. Multiple causes of diagenetic fabric anisotropy in
Gray, M.B., Nickelsen, R.P., 1989. Pedogenic slickensides, indicators of strain and weakly consolidated mud, Nankai accretionary prism, IODP Expedition 316.
deformation processes in redbed sequences of the Appalachian foreland. Ge- J. Struct. Geol. 32, 1887e1898. http://dx.doi.org/10.1016/j.jsg.2010.03.008.
ology 17, 72e75. Mitra, G., Ismat, Z., 2001. Microfracturing associated with reactivated fault zones
Haines, S.H., van der Pluijm, B. a, 2012. Patterns of mineral transformations in clay and shear zones: what can it tell us about deformation history? Geol. Soc. Lond.
gouge, with examples from low-angle normal fault rocks in the western USA. Spec. Publ. 186, 113e140. http://dx.doi.org/10.1144/GSL.SP.2001.186.01.08.
J. Struct. Geol. 43, 2e32. http://dx.doi.org/10.1016/j.jsg.2012.05.004. Moore, J., Byrne, T., 1987. Thickening of fault zones: a mechanism of melange for-
Haines, S.H., Kaproth, B., Marone, C., Saffer, D., van der Pluijm, B., 2013. Shear zones mation in accreting sediments. Geology, 1040e1043.
in clay-rich fault gouge: a laboratory study of fabric development and evolution. Morgenstern, N.R., Tchalenko, J.S., 1967. Microstructural observations on shear
J. Struct. Geol. 51, 206e225. http://dx.doi.org/10.1016/j.jsg.2013.01.002. zones from slips in natural clays. In: Proceedings of the Geotechnical Confer-
Hancock, P.L., Barka, A. a, 1987. Kinematic indicators on active normal faults in ence 1967, pp. 147e152.
Western Turkey. J. Struct. Geol. 9, 573e584. http://dx.doi.org/10.1016/0191- Niemeijer, A., Spiers, C.J., 2005. Influence of phyllosilicates on fault strength in the
8141(87)90142-8. brittleeductile transition: insights from rock analogue experiments. In:
Hemes, S., Desbois, G., Urai, J.L., De Craen, M., Honty, M., 2013. Variations in the Bruhn, D.F., Burlini, L. (Eds.), High-strain Zones: Structure and Physical Prop-
morphology of porosity in the Boom clay formation : insights from 2D high erties, Special Publications. The Geological Society, London, pp. 303e327.
resolution BIB-SEM imaging and mercury injection porosimetry. Netherl. Geo- Niemeijer, A., Spiers, C.J., 2007. A microphysical model for strong velocity weak-
sci./Geol. Mijnb 92, 275e300. ening in phyllosilicate-bearing fault gouges. J. Geophys. Res. 112, B10405. http://
Holland, M., Urai, J.L., van der Zee, W., Stanjek, H., Konstanty, J., 2006. Fault gouge dx.doi.org/10.1029/2007JB005008.
evolution in highly overconsolidated claystones. J. Struct. Geol. 28, 323e332. Niemeijer, A., Marone, C., Elsworth, D., 2008. Healing of simulated fault gouges
http://dx.doi.org/10.1016/j.jsg.2005.10.005. aided by pressure solution: results from rock analogue experiments. J. Geophys.
Holzer, L., Muench, B., Wegmann, M., Gasser, P., Flatt, R.J., 2006. FIB-nano- Res. 113, B04204. http://dx.doi.org/10.1029/2007JB005376.
tomography of particulate systemsdpart I: particle shape and topology of in- Niemeijer, A., Marone, C., Elsworth, D., 2010. Fabric induced weakness of tectonic
terfaces. J. Am. Ceram. Soc. 89, 2577e2585. http://dx.doi.org/10.1111/j.1551- faults. Geophys. Res. Lett. 37 http://dx.doi.org/10.1029/2009GL041689.
2916.2006.00974.x. Nussbaum, C., Bossart, P., 2008. Geology. In: Thury, M., Bossart, P. (Eds.), Mont Terri
Houben, M.E., Desbois, G., Urai, J.L., 2013. Pore morphology and distribution in the Rock Laboratory. Project, Programme 1996 to 2007 and Results. Swiss
Shaly facies of Opalinus clay (Mont Terri, Switzerland): insights from repre- Geological Survey, Wabern, pp. 29e39.
sentative 2D BIBeSEM investigations on mm to nm scale. Appl. Clay Sci. 71, Nussbaum, C., Meier, O., Masset, O., Badertscher, N., 2006. Self-sealing of Fault (SF)
82e97. http://dx.doi.org/10.1016/j.clay.2012.11.006. Experiment Drilling of Resin Impregnated Boreholes Part of Drilling Campaign
Ingram, G.M., Urai, J.L., 1999. Top-seal leakage through faults and fractures: the role of Phase 11 Drilling Data, Drillcore Mapping & Photo Documentation. Mont
of mudrock properties. Geol. Soc. Lond. Spec. Publ. 158, 125e135. http:// Terri Technical Note.
dx.doi.org/10.1144/GSL.SP.1999.158.01.10. Nussbaum, C., Bossart, P., Amann, F., Aubourg, C., 2011. Analysis of tectonic struc-
Ingram, G.M., Urai, J.L., Naylor, M.A., 1997. Sealing processes and top seal assess- tures and excavation induced fractures in the Opalinus clay, Mont Terri un-
ment. Nor. Pet. Soc. Spec. Publ. 7, 165e174. derground rock laboratory (Switzerland). Swiss J. Geosci. 104, 187e210. http://
Ishii, E., 2012. Microstructure and origin of faults in siliceous mudstone at the dx.doi.org/10.1007/s00015-011-0070-4.
Horonobe underground research laboratory site, Japan. J. Struct. Geol. 34, Passchier, C.W., Trouw, R. a. J., 2005. Microtectonics, second ed. Springer-Verlag,
20e29. http://dx.doi.org/10.1016/j.jsg.2011.11.001. Berlin/Heidelberg. http://dx.doi.org/10.1007/3-540-29359-0.
128 B. Laurich et al. / Journal of Structural Geology 67 (2014) 107e128
Pearson, F.J., Acros, D., Bath, D., Boisson, J.Y., Frena ndez, A.M., Ga €bler, H.-E., Urai, J.L., Wong, S.W., 1994. Deformation mechanisms in experimentally deformed
Gaucher, E., Gautschi, A., Griffault, L., Hernan, P., Waber, H.N., 2003. Mont Terri shales. Ann. Geophys. e European Geophysical Union Annual Meeting 12
Project: Geochemistry of Water in the Opalinus Clay Formation at the Mont (Suppl. 1), C98.
Terri Rock Laboratory, Geology. Federal Office for Water and Geology (FOWG), Urai, J.L., Humphreys, F.J., Burrows, S.E., 1980. In-situ studies of the deformation and
Bern, Switzerland. dynamic recrystallization of rhombohedral camphor. J. Mater. Sci. 15,
Petit, J.-P., Laville, E., 1987. Morphology and microstructures of hydroplastic slick- 1231e1240. http://dx.doi.org/10.1007/BF00551812.
ensides in sandstone. Geol. Soc. Lond. Spec. Publ. 29, 107e121. http://dx.doi.org/ Van der Zee, W., Urai, J.L., 2005. Processes of normal fault evolution in a siliciclastic
10.1144/GSL.SP.1987.029.01.10. sequence: a case study from Miri, Sarawak, Malaysia. J. Struct. Geol. 27,
Rutter, E.H., Maddock, R.H., Hall, S.H., White, S.H., 1986. Comparative micro- 2281e2300. http://dx.doi.org/10.1016/j.jsg.2005.07.006.
structures of natural and experimentally produced clay-bearing fault Vannucchi, P., Maltman, A., Bettelli, G., Clennell, B., 2003. On the nature of scaly
gouges. Pure Appl. Geophys. PAGEOPH 124, 3e30. http://dx.doi.org/10.1007/ fabric and scaly clay. J. Struct. Geol. 25, 673e688. http://dx.doi.org/10.1016/
BF00875717. S0191-8141(02)00066-4.
Rutter, E.H., Holdsworth, R.E., Knipe, R.J., 2001. The nature and tectonic significance Vrolijk, P., Sheppard, S.M.F., 1991. Syntectonic carbonate veins from the Barbados
of fault-zone weakening: an introduction. Geol. Soc. Lond. Spec. Publ. 186, 1e11. accretionary prism (ODP Leg 110): record of palaeohydrology. Sedimentology
http://dx.doi.org/10.1144/GSL.SP.2001.186.01.01. 38, 671e690. http://dx.doi.org/10.1111/j.1365-3091.1991.tb01014.x.
Saffer, D.M., Lockner, D. a., McKiernan, A., 2012. Effects of smectite to illite trans- Vrolijk, P., van der Pluijm, B. a, 1999. Clay gouge. J. Struct. Geol. 21, 1039e1048.
formation on the frictional strength and sliding stability of intact marine http://dx.doi.org/10.1016/S0191-8141(99)00103-0.
mudstones. Geophys. Res. Lett. 39, 1e6. http://dx.doi.org/10.1029/ Walsh, J.J., Watterson, J., Bailey, W.R., Childs, C., 1999. Fault relays, bends and
2012GL051761. branch-lines. J. Struct. Geol. 21, 1019e1026. http://dx.doi.org/10.1016/S0191-
Sasseville, C., Clauer, N., Tremblay, A., Pinet, N., 2012. Timing of fault reactivation in 8141(99)00026-7.
the upper crust of the St. Lawrence rift system, Canada, by KeAr dating of illite- Warr, L.N., Cox, S., 2001. Clay mineral transformations and weakening mechanisms
rich fault rocks 1 1 GEOTOP Contribution 2012-0002. Can. J. Earth Sci. 49, along the Alpine Fault, New Zealand. Geol. Soc. Lond. Spec. Publ. 186, 85e101.
637e652. http://dx.doi.org/10.1139/e2012-008. http://dx.doi.org/10.1144/GSL.SP.2001.186.01.06.
Schleicher, A.M., van der Pluijm, B. a., Solum, J.G., Warr, L.N., 2006. Origin and Wenk, H.-R., Voltolini, M., Kern, H., Popp, T., Mazurek, M., 2008. Anisotropy in shale
significance of clay-coated fractures in mudrock fragments of the SAFOD from Mont Terri. Lead. Edge 27, 742e748. http://dx.doi.org/10.1190/1.2944159.
borehole (Parkfield, California). Geophys. Res. Lett. 33, 1e5. http://dx.doi.org/ Will, T., Wilson, C., 1989. Experimentally produced slickenside lineations in pyro-
10.1029/2006GL026505. phyllitic clay. J. Struct. Geol. 11, 657e667.
Sibson, R.H., 1977. Fault rocks and fault mechanisms. J. Geol. Soc. Lond. 133, Wintsch, R., 1998. Strengthening of fault breccia by K-feldspar cementation. In:
191e213. Snoke, A.W., Tullis, J., Todd, V.R. (Eds.), Fault Related Rocks e a Photographic
Sibson, R.H., 2000. Fluid involvement in normal faulting. J. Geodyn. 29, 469e499. Atlas. Princton, pp. 42e43.
http://dx.doi.org/10.1016/S0264-3707(99)00042-3. Yan, Y., 2001. Deformation microfabrics of clay gouge, Lewis Thrust, Canada: a case
Stanley, R.S., 1990. The evolution of mesoscopic imbricate thrust faultsdan example for fault weakening from clay transformation. Geol. Soc. Lond. Spec. Publ. 186,
from the Vermont Foreland, U.S.A. J. Struct. Geol. 12, 227e241. http://dx.doi.org/ 103e112. http://dx.doi.org/10.1144/GSL.SP.2001.186.01.07.
10.1016/0191-8141(90)90007-L. Zulauf, G., Kleinschmidt, G., Oncken, O., 1990. Brittle deformation and graphitic
Tjia, H.D., 1964. Slickensides and fault movements. Geol. Soc. Am. Bull. 75, 683e685. cataclasites in the pilot research well KTB-VB (Oberpfalz, FRG). Geol. Soc. Lond.
Twiss, R.J., Moores, E.M., 1992. Structural Geology. W.H. Freeman, New York. Spec. Publ. 54, 97e103. http://dx.doi.org/10.1144/GSL.SP.1990.054.01.10.