You are on page 1of 17

Tectonophysics 530-531 (2012) 1–17

Contents lists available at SciVerse ScienceDirect

Tectonophysics
journal homepage: www.elsevier.com/locate/tecto

Review Article

A review of brittle compressional wedge models


Susanne J.H. Buiter ⁎
Geological Survey of Norway, Leiv Eirikssons vei 39, 7040 Trondheim, Norway
Physics of Geological Processes, University of Oslo, Sem Selands vei 24, 0316 Oslo, Norway

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

Article history: Fold-and-thrust belts and accretionary wedges form by compression of sedimentary sequences and base-
Received 13 July 2011 ment rocks into forward and backward thrusts, folds, nappes, and duplexes. For over a century, models
Received in revised form 1 December 2011 have been used to investigate the essential characteristics of such brittle wedges. Here I review model studies
Accepted 10 December 2011
of brittle thrust wedges in orthogonal compression, focussing on critical taper theory, analogue and numer-
Available online 23 December 2011
ical techniques for modelling brittle behaviour, and the most commonly investigated variations in wedge
Keywords:
model studies, those in basal dip, basal strength, internal strength, and surface processes. Many model results
Critical taper theory can be placed in the context of critical taper theory, which provides analytical solutions for wedge taper angle
Accretionary wedges and slip line orientations for a homogeneous material on the verge of failure throughout. Dynamic forward
Fold-and-thrust belts models have confirmed critical taper predictions of decreasing surface dip for increasing basal dip, wider
Analogue models wedges with steeper forward thrusts and shallower backward thrusts for decreasing basal strength, narrower
Numerical models wedges and enhanced exhumation for surface erosion, and decreasing surface dip for increasing internal
Plasticity strength. But analogue and numerical models have been able to take these results a step further by investi-
gating the evolution of non-critical thrust wedges and the effects of non-homogeneous materials with décol-
lement layers and strain-weakening shear zones. These results have highlighted the strong impact of
heterogeneous materials on the internal structures of thrust wedges and raise the question whether the com-
plexity of structures that are observed in many natural fold-and-thrust belts requires that lithological layer-
ing, inherited faults, or a wide range of rheologies need to be included in forward models of brittle thrust
wedges.
© 2011 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Critical taper theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3. Modelling brittle deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1. Analogue modelling of brittle deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2. Numerical modelling of brittle deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
4. Models of brittle compressional wedges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4.1. The set-up of wedge models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4.1.1. Model techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4.1.2. Model set-ups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.2. Quantification of model results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.3. Examples of wedge model sensitivities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.3.1. The role of basal dip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.3.2. The role of basal strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.3.3. The role of surface processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4.3.4. The role of internal strength variations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
5. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

⁎ Geological Survey of Norway, Leiv Eirikssons vei 39, 7040 Trondheim, Norway. Tel.: + 47 73904265.
E-mail address: susanne.buiter@ngu.no.

0040-1951/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2011.12.018
2 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

1. Introduction FTBs has come from their association with hydrocarbons, such as in
the Zagros, Andes and Caucasus (Cooper, 2007), and for AWs from
Compressional, thrust or Coulomb wedges are general terms that their association with sources of large earthquakes and related tsu-
describe fold-and-thrust belts and accretionary wedges with perhaps namis. These case studies have been complemented by analytical, nu-
a slight model flavour. Fold-and-thrust belts (FTBs) form in the fore- merical and analogue studies of compressional wedges. The analytical
land of an orogen and include sedimentary sequences, which often critical taper theory of Coulomb wedges has been very successful in
derive from a former passive margin, and a variable amount of base- describing the geometry of wedges in response to accretion, changes
ment rocks. The latter determines whether an FTB is thin-skinned, in internal strength, and variations in basal strength (Dahlen, 1984;
mainly involving sediments and only shallow levels of basement, or Dahlen et al., 1984; Davis et al., 1983). However, this theory only pro-
thick-skinned, where thrusts and folds affect the entire upper crust vides a ‘static’ view of wedges that assumes constant stress conditions
and possibly the lower crust (Coward, 1983; Pfiffner, 2006; Rodgers, and a wedge at the verge of failure everywhere. Insight into the dy-
1949). Examples can be found, among others, in the Appalachians, namic evolution of wedges and their non-critical behaviour can be
the Canadian Rocky Mountains, the European Alps, and Taiwan (e.g., obtained by analogue and numerical model studies.
Davis et al., 1983; Pfiffner, 2006; Poblet and Lisle, 2011; Price, 1981) Analogue, or laboratory, models are scaled experiments that use
(Fig. 1a). FTBs are typically associated with ocean–continent conver- low strength materials such as sand, clay, wax and silicone putty to
gence, terrane accretion, continent–continent collision, or intraconti- simulate tectonic processes (e.g. Cadell, 1890; Hubbert, 1937; Koyi
nental shortening. They display top-to-the foreland thrusts (which and Mancktelow, 2001; Willis, 1893). The models are intrinsically 3-
may bound nappes or form duplex structures), fault-related folds, de- D and allow direct observations of the experiment, while deformation
tachment zones, triangle zones, back-thrusts, or out-of-sequence may be quantified using image analysis (e.g., Particle Image Veloci-
thrusting (Butler, 1982; Pfiffner, 2006; Poblet and Lisle, 2011). Accre- metry), laser scanning (miniature laser altimetry), and volumetric
tionary wedges (AWs) can be seen as a subset of fold-and-thrust scanning using computerised X-ray tomography (Adam et al., 2005;
belts. They are formed by continent-derived or oceanic sediments off- Colletta et al., 1991). However, although experiments can be repeated
scraped from the subducting plate at an active margin and are usually with similar results, exact reproducibility does not exist. Moreover,
submarine. Sediments can be incorporated into an AW by frontal ac- laboratories that do not have access to tomographic techniques can
cretion near the toe of the wedge or through basal accretion after only monitor the outside boundary of models, with observation of in-
being taken partway down the subduction channel (Glodny et al., ternal structures limited to cutting the model open at the end of the
2005; Gutscher et al., 1998a; Von Huene and Scholl, 1991). Examples experiment. Numerical experiments are exactly reproducible (unless
of accretionary wedges can be found, among others, in Nankai, Mak- they are seeded by random noise). They allow great flexibility in
ran, the Lesser Antilles, and Barbados (Moore and Silver, 1987; choice of geometries and material properties, and can quantify a
Morgan and Karig, 1995; Morley et al., 2011; Westbrook et al., large range of parameters directly, such as, stress, strain, velocity,
1982; Westbrook et al., 1988) (Fig. 1b). AWs and thin-skinned FTBs and energy dissipation. Their resolution in 3-D may still be limited,
show considerable variability owing to the variations in their tectonic but this is improving rapidly (e.g., Braun and Yamato, 2010; Gerya,
setting, evolution and mechanical properties, but a few common ob- 2010; Popov and Sobolev, 2008).
servations are (e.g., Chapple, 1978): (1) a relatively weak basal de- In this paper, I will review numerical and analogue studies of
tachment or décollement which dips towards the interior of the compressional wedges focussing on the roles of parameters that
orogen and below which there is little deformation, (2) large horizon- are often investigated: basal dip, basal strength, surface processes,
tal shortening above the basal detachment, (3) thrusts which dip to- and internal strength variations. I have limited my review to wedges
wards the hinterland (i.e., the orogen for FTBs or the overriding plate of brittle materials that are shortened in orthogonal compression
for AWs) and which systematically become younger towards the and mainly respond by forming brittle structures. Examples of vis-
foreland, and (4) a characteristic wedge shape, which tapers towards cous wedges can be found in, for example, Emerman and Turcotte
the foreland. (1983), Rossetti et al. (2000), Medvedev (2002), and Luján et al.
Fold-and-thrust belts and accretionary wedges are relatively ac- (2010), while McClay et al. (2004) and Del Castello et al. (2005)
cessible and many geological and geophysical observations have show examples of oblique wedges. In the next sections I first de-
helped to obtain a good understanding of their structures. This pro- scribe the analytical theory of critical tapers, followed by a discus-
vides direct input into studies of topography evolution, feedback rela- sion of numerical and analogue techniques for modelling brittle
tions between topography and climate, landslide risk, and active behaviour, before finally reviewing dynamic models of brittle
margin processes. An additional motivation for detailed studies of wedges.

a)
W
E
0
time (s)

16
0 5 10 km

b) W
sediments
reflection time (s)

E
8 wedge detachment
sediments
10 0 5 10 km

Fig. 1. a) Cross-section through the Canadian Rocky Mountains fold-and-thrust belt, redrawn from Bally et al. (1966). b) Cross-section through the Lesser Antilles acrretionary
wedge, redrawn from Westbrook et al. (1982).
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 3

2. Critical taper theory of brittle material which is on the verge of failure everywhere (i.e.,
static stress conditions); b) Rock strength is time-invariable and de-
In a series of influential papers in the early 1980s it was shown termined by the Coulomb criterion:
that insight into the overall structural evolution of thin-skinned
fold-and-thrust belts and accretionary wedges can be obtained by jτj ¼ C þ μ ð1−λÞσ n ¼ C þ tan ϕð1−λÞσ n ð1Þ
considering them analogous to wedges of snow in front of a moving
bulldozer (Dahlen, 1984; Dahlen et al., 1984; Davis et al., 1983) τ is shear stress, σn normal stress, C cohesion, μ friction coefficient,
(Fig. 2a). The material in front of the bulldozer will deform until it μ = tan ϕ, ϕ angle of internal friction, and λ the pore fluid factor de-
reaches a so-called critical taper. If no further material is encountered, fined as the ratio of pore fluid pressure to normal stress (Hubbert
the wedge will slide stably along the base, without internal deforma- and Rubey, 1959); c) The setting is 2-D plane strain and sidewall fric-
tion. If new material is encountered, the wedge will grow self- tion is not considered; d) Material properties are constant in time, ho-
similarly at its critical taper value. Critical taper theory is successful mogeneous in space, and isotropic; e) The base is cohesionless; and
at explaining the overall geometry of wedges, as defined by their sur- f) The strength of the base cannot exceed the strength of the interior.
face slope and basal dip, the orientation of slip lines, and the depen- Here compression is taken as positive. The critical taper equation can
dence of these properties on internal and basal strength. Critical be derived from the equations of stress equilibrium, the Coulomb
taper theory uses the following assumptions: a) The wedge consists criterion, and the boundary conditions of a shear stress free top and

a) b)
ρw
D α ψ0 x

α+β z
θf
ψb θb

σ1 β ρ
σ3
c)
30o d)
w2 critical taper
20o λ=0
unstable
surface slope α

10o field o
21.8
0 o w1
stable field
-10o w2
o
-20
critical taper
-30o λ = 0.3
w1 4.6
o

-10o 0o 10o 20o 30o 40o 50o 60o 70o 80o 90o
basal slope β
e) f)
40 o 40o
angle of repose 40o
Variation in φb Variation in φ
surface slope α
surface slope α

20 o 20o 34o
32 o
17 o
2 o

o
0 0o
18o

-20 o -20o
φ = 34o φb = 17o
λ=0 λ=0
-40 o -40o
-20 o
0 o
20 o
40o
60 o
80 o
-20o 0o 20o 40o 60o 80o
basal slope β basal slope β

Fig. 2. a) Cartoon illustrating stable sliding of a snow wedge before an accretion event (top) and an actively accreting snow wedge deforming to reach its critical taper (bottom),
After Dahlen and Suppe (1988). b) Cross-section through a wedge with a critical taper angle α + β. ψb is the angle between the maximum compressive stress σ1 and the base of the
wedge, ψ0 the angle between σ1 and the top of the wedge, θf the dip angle of forward thrusts with the base of the wedge, θb the dip angle of backward thrusts with the base of the
wedge, ρ wedge density, ρw water density, and D local water depth. After Dahlen and Suppe (1988). c) Wedge stability field for ϕ = 34°, ϕb = 17°, and λ = λb = 0.3 (grey field) or
λ = λb = 0 (dotted lines). d) Two examples of critical wedges with a horizontal basal slope (values as in c) with λ = λb = 0.3). Wedge w1 deforms by reverse faulting, whereas the
wedge w2 deforms by normal faulting. e) An increase in basal friction angle ϕb decreases the wedge stability field, leading to an increase in the lower critical taper angle and a
decrease in the upper taper angle. f) An increase in the angle of internal friction ϕ increases the wedge stability field.
4 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

a prescribed friction on the base (Barcilon, 1987; Dahlen, 1984; the base (θf) and backward thrusts 35.3° (θb) (w1 in Fig. 2d). This
Dahlen et al., 1984; Davis et al., 1983; Dahlen, 1990): wedge has a sharp taper angle and fails by thrusting. The second crit-
ical surface slope (and taper angle) is 21.8°. For this wedge,
α þ β ¼ ψb −ψ0 ð2Þ ψb = 65.7°, ψ0 = 43.9°, θf = − 37.7°, and θb = 93.7° (w2 in Fig. 2d).
This wedge fails by normal faulting. Wedges that are built from ini-
α is the surface slope, β the basal slope and their sum the taper angle,
tially quasi-horizontal strata will in most cases first achieve the
which should be positive (Fig. 2b). ψb and ψ0 are the angles between
lower critical taper angle (e.g., w1 in Fig. 2c and d) and might then
the maximum principal stress σ1 and the base and top of the wedge,
further adjust their geometry by building tapers consistent with this
respectively. For a cohesionless wedge:
critical taper solution.
! " Critical taper theory can be used to understand the effects of var-
1 sinϕ′b 1 ′
ψb ¼ arcsin − ϕb ð3Þ iations in basal strength, basal dip, internal wedge strength, erosion,
2 sinϕ 2
and sedimentation. An increase in basal friction ϕb will decrease the
! " stability field (Fig. 2e), by narrowing it vertically and widening it hor-
1 sinα ′ 1 ′
ψ0 ¼ arcsin − α ð4Þ izontally. This means that a critically-tapered wedge can find itself in
2 sinϕ 2
the unstable field and will need to adjust its taper angle. Dahlen
! " (1984) shows that this can be accompanied by a change in tectonic
1−λb
tan ϕ′b ¼ tan ϕb ð5Þ style within the wedge or even the need to form a new basal décolle-
1−λ
ment. For wedges at the lower critical taper value (e.g., the w1 wedge
!! " " in Fig. 2d), an increase in basal friction causes an increase in the angle
′ 1−ρw =ρ
α ¼ arctan tan α ð6Þ that the maximum compressive stress makes with the base of the
1−λ
wedge (ψb), an increase in taper angle, shallower forward thrusts,
ϕb is the angle of basal friction, ρ density, ρw density of water, λ the and steeper backward thrusts. An increase in internal wedge strength
generalised pore fluid factor in the wedge, and λb the generalised ϕ increases the stability field (Fig. 2f), by widening it vertically, there-
basal pore fluid factor: by decreasing the lower critical taper angle (and increasing the upper
one). An increase in internal wedge strength has a similar effect as a
P−ρw gD relative decrease in basal strength, i.e., a vertical widening of the sta-
λ¼ ð7Þ
σ z −ρw gD bility field (Fig. 2e, f). If material is removed from the surface by ero-
sion in a pattern that changes taper angle (e.g., height-dependent
P b −ρw gD erosion rate), wedges at the lower critical taper angle will need to de-
λb ¼ ð8Þ
σ z −ρw gD form to maintain their critical taper. Dahlen and Suppe (1988) (see
also Dahlen and Barr, 1989) show that wedges achieve a flux
P is pressure, Pb pore fluid pressure on the basal detachment, σz ver- steady-state (Willett and Brandon, 2002) if erosion balances the ac-
tical stress, g gravitational acceleration, and D local waterdepth for cretion of new material at the toe of the wedge. In that case, relatively
submarine wedges. For subaerial wedges, the above equations simpli- high erosion rates (∼5–10 mm yr −1) are expected to lead to narrow
fy by D = ρw = 0. In the coordinate system of Dahlen (1984), σz is ap- mountain belts (50–200 km), such as in Taiwan and New Zealand,
proximately equal to σn for small taper angles. Wang et al. (2006) whereas relatively low erosion rates (∼0.5 mm yr −1) might explain
show how the expression for λb can be adjusted to avoid a stress dis- wide belts (>500 km), such as the Central Andean Altiplano.
continuity at the base of the wedge for taper angles above ca. 10°. The Critical taper theory captures an average state of a fold-and-thrust
angle of thrust faults with
# the
$ maximum compressive stress σ1 in a belt and has a number of limitations: First, it is assumed that wedges
Coulomb material is % π4 − ϕ2 (Coulomb, 1773). The dip angle of for- are at the verge of failure everywhere and the theory can, therefore,
ward thrusts, θf, and backward thrusts, θb, relative to the base of the not account for wedges that are partly or transiently not at failure
wedge is then (Fig. 2b): (Lohrmann et al., 2003; Simpson, 2011; Stockmal et al., 2007; Wang
and Hu, 2006). Second, materials are assumed homogeneous and iso-
π ϕ
θf ¼ − −ψb ð9Þ tropic and memory effects through, for example, reactivation of
4 2
strain-weakened fault planes are not considered. Third, the stress
π ϕ state for wedges in the stable field is not defined, but could be ap-
θb ¼ − þ ψb ð10Þ proximated by assuming linear elasticity (Wang and Hu, 2006).
4 2
Fourth, critical taper theory explains the present-day state and overall
Critical cohesionless wedges are self-similar and have constant geometry of wedges, but does not show how the structures evolved
taper angle and stress orientation throughout their interior. This is to this state. Thrust wedges need finite deformation, averaging over
still the case for cohesive wedges which have a cohesion that in- individual deformation events, to reach a long-term state. Wang and
creases linearly with depth (Zhao et al., 1986). Wedges with a con- Hu (2006) suggest that the actively deforming, seaward part of accre-
stant cohesion have, however, a slightly concave upper surface tionary wedges cycles through critical and stable states triggered by
(Dahlen et al., 1984). An example of critical taper values for cohesion- large earthquakes and interseismic periods, respectively. The surface
less rock and sand wedges is shown in Fig. 2c. It can be seen that two slope of this outer wedge is determined by the peak basal stress in
branches exist, with a lower and an upper surface ‘critical’ slope value large earthquakes and will be established over many earthquake cy-
for any given basal dip. The maximum surface slope for dry (λ = 0) cles. In addition, the exact cohesionless solution of Dahlen (1984) re-
and subaerial wedges is the angle of repose, αmax = ϕ (Dahlen, laxes the constraint that σ1 is parallel to the wedge surface (this is
1984). Wedges that plot in the area between the two branches are possible as σ1–σ3 vanishes along the surface). For this reason, slip
stable, whereas wedges in the fields outside will deform in order to lines are planar surfaces in the exact solution (Fig. 3a). If σ1 is parallel
reach the critical taper value. Many analogue and numerical thrust to the wedge surface, however, faults become listric with dip
wedge models use a horizontal base, β = 0°. I here illustrate an exam- angle increasing upwards (Dahlen et al., 1984; Davis et al., 1983)
ple for a wet subaerial wedge with ϕ = 34° (e.g., (Byerlee, 1978), λ = (see also Hafner, 1951; Hubbert, 1951) (Fig. 3b). The degree to
λb = 0.3, and ϕb = 17° (taken as ϕ/2) (Fig. 2c). For β = 0°, α is 4.6° and which faults are listric depends on basal friction. A very weak base
the critical taper angle (α + β) is therefore also 4.6°. ψb = 7.3° and (ϕb ≪ ϕ) leads to almost planar surfaces, whereas high basal friction
ψ0 = 2.6° (Eqs. 3 and 4). Forward thrusts make a 20.7° angle with (ϕb→ ϕ) results in listric faults (Fig. 3b, c). Despite the above
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 5

σ1
a) Dahlen (1984) softening occurs until an often stable dynamic strength is reached
(Han and Drescher, 1993; Lohrmann et al., 2003; Panien et al.,
ψ0
2006; Schanz and Vermeer, 1996) (Fig. 4a). This behaviour resembles
the stress–strain curves of rocks in compression tests (Jaeger et al.,
2007) and, together with common friction coefficients on the order
α = 17.8o
of ca. 0.6–0.7, is the main argument for macroscopic similarity be-
ψb tween the mechanics of sandbox models and Earth's crust. Initial
compaction of a granular sample under the imposed shear deforma-
σ1 tion is followed by dilation with formation of shear bands close to
the peak strength and maximum dilation rate. The dynamic stable
b) Davis et al. (1983) - high basal friction σ1
state is associated with decreased decompaction rates.
The frictional properties of granular materials used in analogue
modelling experiments have large variations as shown by a survey
of sands and glassbeads used in modelling laboratories around the
α = 17.8o world by Klinkmüller (2011). Under identical laboratory measure-
ment conditions, this study found peak angles of internal friction
ψb
which varied between 26° and 41°, and dynamic stable angles of in-
ternal friction varying between 22° and 35°. As the angles of internal
σ1
friction and strain softening impact shear zone dip angles and strain
localisation, these variations illustrate the importance of accurate
c) Davis et al. (1983) - low basal friction measurements of the properties of granular materials used in an ex-
σ1 periment. In addition, the peak angle of internal friction has been
α = 3.4o
shown to vary with material density, grain shape, and handling tech-
σ1 nique (Krantz, 1991; Lohrmann et al., 2003; Mair et al., 2002; Schanz
and Vermeer, 1996; Schellart, 2000). Sifted or sprinkled granular ma-
Fig. 3. a) The exact cohesionless wedge solution of Dahlen (1984) has a constant stress terials have a higher angle of peak friction and show more strain-
field throughout the wedge, relaxing the constraint that σ1 should be parallel to the softening than poured materials. This is because sifting or sprinkling
surface. This results in planar potential fault surfaces (sliplines). This example has will result in a denser packing and, therefore, higher density than
Φ = 30∘, Φb = 29.9∘ and β = 0∘. b) If σ1 is parallel to the upper surface, a wedge with
high basal friction will have curved fault surfaces. Φ = 30∘ and Φb = 29.9∘. c) Wedges
pouring (Krantz, 1991). A denser packing allows less compaction
with low basal friction have an almost constant direction for σ1 throughout the when the material is initially loaded, but well-compacted sifted
wedge, also in case σ1 is constrained to be parallel to the upper surface. This results sands show more dilation associated with the formation of a shear
in nearly planar potential fault surfaces. Φ = 30∘ and Φb = 10∘. band, and a larger degree of strain-softening (Lohrmann et al.,
2003) (Fig. 4a). Poured granular materials in contrast show substan-
limitations, the analytical solutions are powerful and fast tools to tial initial compaction with limited decompaction in the shear band.
grasp the first-order behaviour of compressional wedges. As a result, the dynamic stable angle of friction seems not to depend
on the handling technique, while the peak angle of friction does
3. Modelling brittle deformation (Lohrmann et al., 2003).
Cohesion is generally obtained from shear-tests by extrapolation
3.1. Analogue modelling of brittle deformation of the measurements towards zero normal stress using a linear re-
gression model. This may lead to high cohesion values. However,
Analogue models of brittle deformation typically use anorganic or some studies have found that the failure envelope has a convex out-
artificial granular materials, such as quartz sand, corundum sand, ward shape at low normal stresses (below ca. 400 Pa) and a cohesion
microbeads, microspheres, but also a range of organic materials, of 0 Pa at zero normal stress has been inferred for at least some sands
such as, wheat flour or crushed walnut shells (e.g. Cruz et al., 2008; (Richard and Krantz, 1991; Schellart, 2000) (Fig. 4b). Alternatively,
Hubbert, 1951; Lohrmann et al., 2003; Panien et al., 2006; Rossi and Mourgues and Cobbold (2003) show that sidewall friction between
Storti, 2003; Schreurs et al., 2006). The widespread use of sand as a the granular material and the container can cause an overestimation
rock analogue is reflected in the term ‘sandbox’ model for this type of the cohesion (and an underestimation of the coefficient of friction).
of experimental approach. Also clay has been shown to successfully The true cohesion might be lower than reported in many shear-tests.
reproduce brittle structures of the Earth's crust at laboratory scales Panien et al. (2006) found that the dip angle of extensional shear
(e.g., Cloos, 1955; Withjack and Jamison, 1986). Eisenstadt and Sims zones in quartz sand close to the surface would agree with a convex
(2005) show that dry sand and wet clay produce similar deformation outward yield envelope. However, because principal stresses are
patterns, but with differences in faulting structures that might be at- higher in contraction (the minimum principal stress, σ3, being equal
tributed to the higher cohesion and smaller grain size of wet clays rel- to the overburden stress), the exact shape of the yield envelope at
ative to dry sands. The frictional properties of analogue brittle low stresses will be less likely to strongly affect shear zone dip angles
materials at low stresses equivalent to laboratory conditions (i.e., a in compressional-type experiments. In contraction, therefore, the ex-
few kPa) can be inferred from measurements with a ring-shear tester trapolated cohesion probably captures the relevant part of the failure
(Panien et al., 2006; Schulze, 1994) or a Hubbert-type shear box envelope adequately.
(Hubbert, 1951; Krantz, 1991; Schellart, 2000). Both approaches de- The width of shear bands that develop after failure depends on
termine the friction coefficient and cohesion from measurements of grain size and is observed to be between 10 and 16 times the grain di-
material strength as a function of normal load, as the slope and axis ameter (Hermann, 2001; Mühlhaus and Vardoulakis, 1987; Panien
intercept, respectively, of the data in Mohr space. In general, mea- et al., 2006). In granular media, the dip angle of shear bands with
surements with a ring-shear tester are faster and may be more pre- the maximum compressive stress σ1 has been shown to vary between
cise (Lohrmann et al., 2003). Upon loading, most granular materials the Roscoe (1970) angle:
that are typically used in sandbox-type modelling show limited elas-
tic deformation (since their bulk moduli are on the order of GPa), fol-
lowed by strain-hardening before failure at peak strength, after which θR ¼ 45−ψ=2 ð11Þ
6 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

a) b)
shear stress

σ3
σ3
peak strength
τ
compacted σn 90 - θ σ1
softening
dynamic stable strength
loose θ
σ1
τ

variation of sample thickness


(compacted sand)
C0

tanφ
shear strain σ3 σ1 σn

Fig. 4. a) Schematic illustration of stress–strain curves for dense and loose granular materials with a compaction–decompaction curve for dense (compacted) materials superposed.
After Lohrmann et al. (2003), Panien et al. (2006) and Vermeer (1990). b) Mohr diagram for a sand-like material in contraction. σ3 is vertical and corresponds to the overburden
stress, σ1 is the maximum compressive stress, σn normal stress, τ shear stress, Φ angle of internal friction, C0 is cohesion obtained by extrapolating the Mohr yield envelope to zero
normal stress. Some sands may have zero cohesion at zero normal stress, as indicated by the dashed line.

the Coulomb (1773) angle: corresponds to several grains in an analogue model and particles
should, therefore, not be equated with grains. In addition, the macro-
θC ¼ 45−ϕ=2 ð12Þ scopic frictional properties have to be determined through calibration
experiments on the particle assembly. This latter limitation is over-
and the intermediate Arthur et al. (1977) angle (see also Vardoulakis,
come in the stress-based DEM from Egholm (2007) where Mohr–
1980):
Coulomb failure is prescribed at the particle level. DEM models of
θA ¼ 45−ðϕ þ ψÞ=4 ð13Þ shear bands have illustrated how the nature of deformation within
shear zones depends on sorting, particle size, and particle rotations
ψ is the angle of dilation (Roscoe, 1970): (Antonellini and Pollard, 1995; Bardet and Proubet, 1991; Morgan
and Boettcher, 1999). The friction coefficient within shear zones in-
ε_ 1 þ ε_ 3 creases with angularity of the grains, as is shown in the 3-D DEM
sin ψ ¼ ð14Þ
ε_ 3 −ε_ 1 models with breakable grains of Abe and Mair (2009). Similar to
shear zone width depending on grain size in laboratory experiments,
Shear bands at the Coulomb angle follow the stress characteristics shear zone width in DEM models also depends on particle size. Unfor-
and coincide with maximum shear stress to normal stress ratio. tunately, it seems that the dip angle of shear bands in DEM models
Roscoe shear bands coincide with velocity characteristics (velocity has so far not been systematically addressed. It is, therefore, not
discontinuities). Coulomb angles tend to occur in fine sands, whereas known whether in DEM models a finer particle assembly would result
Roscoe angles are observed in coarse sands, but all shear band incli- in Coulomb dip angles (Eq. 12) and coarser assemblies in Roscoe dip
nations between the Coulomb and Roscoe solutions are possible angles (Eq. 11), similar to what has been shown for sands in the
(Vermeer, 1990). These shear band dip angle solutions are valid for laboratory.
initial stages, as later deformation may rotate shear bands to a differ- Continuum numerical methods often use a smooth version of the
ent angle. In addition, Hubbert (1951) already pointed out that shear (angular) Mohr–Coulomb failure criterion, in the form of the Drucker–
zone dip angle may differ between the interior of an experiment and Prager criterion which can be written in coordinate-independent stress
the sidewalls, due to friction along the sidewalls. invariants as:

3.2. Numerical modelling of brittle deformation σ e ¼ P sin ϕ þ C cos ϕ ð15Þ

As thrusts are a characteristic feature of wedges, numerical σe is the effective stress


# as determined
$1 by the second invariant of the
1 2
models of deforming wedges need to be able to form shear zones fol- stress tensor, σe = σ σ
2 ij ij , and dynamic pressure P is the mean
lowing a failure criterion. This makes techniques that prescribe fault stress. Parts of the model that are not (yet) at failure behave viscous
locations a priori (e.g., through node splitting, Melosh and Williams, or viscoelastic. The behaviour of material that is at yield is specified
1989) less suitable for these type of problems. Shear bands have with a plastic flow-rule that can be associated (ϕ = ψ) or non-
been shown to form successfully in the particle-based distinct- associated (ϕ ≠ ψ). Shear bands in continuum models depend on ele-
element method (DEM) and in continuum methods, such as the ment size or node spacing, but have no intrinsic length scale. This
finite-element method (FEM) and finite-difference method (FDM). means that shear bands can become very narrow for very fine grids.
DEM techniques simulate a material as an assembly of particles and Several solutions that introduce a length scale to shear bands have
could, therefore, be considered intrinsically suited to model granular been proposed (de Borst and Sluys, 1991; Mühlhaus and Vardoulakis,
materials, such as sand (Cundall and Strack, 1979). Dilation effects 1987; Wang et al., 1996), but these techniques have not been investi-
(Fig. 4a) are, for example, included implicitly. However, usually the gated systematically for geodynamic problems. The dip angle of shear
resolution of DEM models is such that one particle in the model zones in continuum models has been shown to vary between the
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 7

theoretically admissible Roscoe, Arthur and Coulomb dip angles (Eqs. models. Model details are in Table 1. Fig. 5b shows that a fine mesh
11–13) and has in fact been a topic of debate for some time (Kaus, and a well-resolved heterogeneity to initiate the shear bands result
2010; Lemiale et al., 2008; Moresi and Mühlhaus, 2006; Popov and in a shear zone dip angle which is very close to the Coulomb dip
Sobolev, 2008; Vermeer, 1990). Lemiale et al. (2008) illustrate that a angle. A coarser mesh and a smaller inclusion lead to steeper dip an-
fine numerical mesh is required to achieve the Coulomb dip angle. In gles, which follows the results of Lemiale et al. (2008) and Kaus
addition, Kaus (2010) shows that the heterogeneity, which is used to (2010). A Roscoe dip angle results for a rather coarse mesh (Fig. 5c),
initiate a shear band, needs to be well resolved (with 5–20 elements) whereas a small, 1 × 1 element inclusion gives an Arthur dip angle
for Coulomb angles. This implies that shear zones that initiate from (Fig. 5d). The use of lithostatic pressure instead of dynamic pressure
small heterogeneities or numerical noise will have orientations close in the Drucker–Prager failure criterion (Eq. 15) results in Roscoe
to the Arthur or Roscoe angle. The rheology of unyielded parts was shear zone dip angles (Fig. 5e). In continuum numerical models, a
found to be less important, though viscoelastic–plastic models in com- Coulomb dip angle for shear zones therefore requires a fine mesh,
pression obtain higher strain rates inside the shear bands than visco- heterogeneities that are well resolved, and the use of dynamic pres-
plastic models (Kaus, 2010). sure (mean stress) instead of lithostatic pressure in the failure
I illustrate the sensitivity of numerical shear zone dip angle to criterion.
mesh size and inclusion size in Fig. 5 for four simple compression
4. Models of brittle compressional wedges
a) Model set-up
free surface 4.1. The set-up of wedge models

60 km 4.1.1. Model techniques


2x10-11 m/s 15 km φ = 30o 15o -2x10-11 m/s
Models of brittle compressional wedges usually simulate a sedi-
C = 10 MPa
seed
mentary/upper crustal domain which is some kilometres thick and
several tens of kilometres wide. Analogue models need therefore to
scale up from the cm-scale of the lab to the km-scale of a natural pro-
b) Reference model totype. Hubbert (1937) discusses the geometric, kinematic and dy-
namic similarity relationships that need to be observed when
scaling to natural dimensions. In a brittle regime, stresses can be
scaled by requiring that the ratio of frictional strength to lithostatic
pressure is the same for model and nature:
! " ! "
σe σe
¼ ð16Þ
c) Coarse grid 5 km ρgh model ρgh nature

σe is the effective stress, ρ density, g gravitational acceleration, and h


a length scale. For example, if a length of 1 cm in a model corresponds
to 1 km in nature, angle of internal friction and density are similar for
the model and the natural prototype, and the experiment is run at
normal gravity, 10 Pa in the model would correspond to ca. 1 MPa
d) Small seed in nature. Numerical models can be dimensionless, at km-scale, or
at cm-scale for direct comparisons to laboratory models. Analogue
models are built by sprinkling, sifting or pouring of a granular mate-
rial into the modelling apparatus. Density, angle of internal friction
and the degree of strain-softening are not only determined by
the choice of material, but also by this handling technique (see
Section 3.1). Variations in basal friction can be obtained by using dif-
e) Lithostatic pressure ferent materials for the base of the apparatus, such as wood, a plastic
sheet or sandpaper, or by inserting a layer of different properties (e.g.,
microbeads or viscous silicone putty) directly above the base. Numer-
ical models of brittle deformation solve for conservation of mass and
strain-rate inv (1/s)
momentum, most commonly under the assumption of quasi steady-
0 5e-14 1e-14 state, which means that acceleration terms are neglected. Conserva-
tion of mass is given by:
Fig. 5. Numerical sensitivity of shear zone dip angles in compression. a) The model is
60 km wide by 15 km high and is shortened by a velocity of 2 × 10−11 m s−1 at each ∂ρ
þ ∇⋅ðρuÞ ¼ 0 ð17Þ
side. The material is incompressible (ψ = 0°) with an angle of internal friction ϕ of ∂t
30° softening to 15° and a cohesion of 10 MPa. The inclusion is viscous with ϕ = 0°. Fur-
ther input values are in Table 1. The Coulomb dip angle for this model is 30°, the Arthur u is the velocity. An assumption that is often used is of incompressi-
dip angle 37.5°, and the Roscoe dip angle 45°. The results are obtained with SULEC (de- bility, in which case ∂ρ∂t
= 0, and the equation reduces to ∇ ⋅ u = 0.
veloped by Susanne Buiter and Susan Ellis) which is an arbitrary Langrangian Eulerian
This implies that dilation ψ = 0, which affects the dip angle of Roscoe
finite-element code. The quadrilateral elements are linear in velocity and constant in
pressure. Pressure is solved with an iterative penalty formulation (Cuvelier et al., and Arthur type shear bands (Eqs. 11 and 13). The equilibrium equa-
1986). b) The ‘reference’ set-up has a resolution of 480 × 120 elements (horizontal × - tion for slow deformation is:
vertical), an inclusion size of 500 × 500 m (4 × 4 elements), and uses dynamic pressure
in the Drucker–Prager failure criterion (Eq. 15). The shear-zone dip angle after 25 time ′
∇ ⋅ σ þ ∇P−ρg ¼ 0 ð18Þ
steps is 31° which is very close to the Coulomb dip angle of 30°. c) A coarse grid of
100 × 25 elements results in a Roscoe dip angle of 45°. d) A smaller inclusion size of
125 × 125 m (1 × 1 elements) results in a dip angle of 37° which is as good as identical
σ ′ is the deviatoric stress and P is dynamic pressure. These equations
to the Arthur dip angle. e) The use of lithostatic pressure instead of dynamic pressure in can be solved with different numerical techniques (e.g., finite ele-
the failure criterion results in shear zones with a Roscoe dip angle (45°). ment, finite difference, or distinct element techniques) and with
8 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

Table 1 2001; Simpson, 2009; Smit et al., 2003; Teixell and Koyi, 2003). The
Model values for the shear zone experiment of Fig. 5. wedge builds in front of the mobile wall and especially the first stages
Parameter Value of deformation will be sensitive to the shape of the strong wall, to the
presence or absence of an exit slot beneath the mobile wall and, for
Width 60 km
Height 15 km numerical models, to the accuracy of the pressure solution at
Element size 125 × 125 m the basal corner velocity discontinuity. A pre-existing wedge (e.g.,
Angle of internal friction ϕ 30° → 15° Fig. 7a) may help to form the first thrusts away from the mobile
Cohesion C 10 MPa
wall, though results point to varying success (Buiter et al., 2006;
Density 2700 kg m−3
Background viscosity 1025 Pa s Cubas et al., 2010; Schreurs et al., 2006). The wall acts as a rigid back-
Penalty (compressibility) factor 1032 Pa s stop and plays the role of an analogue to the bulldozer in the critical
Inclusion width 500 m taper theory (Fig. 2a). It represents material that is stronger than
Inclusion height 500 m the incoming sediments and could, for example, be identified with a
Inclusion angle of internal friction 0°
crystalline part of the overriding plate or with sediments that were
Inclusion viscosity 1020 Pas
Gravitational acceleration 10 m s−2 accreted in earlier phases and that are (partly) lithified (Byrne et al.,
Maximum yield stress 1 kPa 1993). The impact of the vertical backstop has been investigated in
Maximum yield stress 1000 MPa experiments that vary the backstop shape (Persson, 2001) or that at-
Strain weakening interval 0–0.1
tach plates of various thickness to the mobile wall thus placing the
Velocity V (each side) 2 × 10−11 m s−1
Time step 10 kyr basal velocity discontinuity outwards (Haq and Davis, 2008; Smart
and Couzens-Schultz, 2001). In the latter case, thrusts will form
away from the mobile wall and material can be (back-)thrust over
different resolution of the discretisation of the spatial domain. This the indentor (depending on the indentor height), forming a two-
will impact shear band width and dip angle (Section 3.2). Also, differ- sided or double-vergent wedge. As Koons (1990) pointed out, this
ent techniques can be used to apply boundary friction, for example, may be a more realistic condition than an indentor that is much
incorporating frictional layers explicitly within the model domain or higher than the thickness of the incoming material. For thin sheets at-
iterating to keep boundary shear stresses at yield (e.g., Buiter et al., tached to the mobile wall, the set-up becomes similar to a S-point ex-
2006). The results of such boundary friction techniques have so far periment (see below).
not been compared directly. In conveyor-type (pull-type) experiments, material is dragged
towards the backstop (Fig. 6b) (e.g. Agarwal and Agrawal, 2002;
4.1.2. Model set-ups Byrne et al., 1993; Davis et al., 1983; Gutscher et al., 1998b;
Fig. 6 shows examples of set-ups that are commonly used in brittle Konstantinovskaia and Malavieille, 2005; Kukowski et al., 2002;
wedge modelling. Push-type experiments (Fig. 6a) achieve shorten- Lallemand et al., 1994; Stockmal et al., 2007). The model domain in
ing by moving a mobile wall into the deformable model domain. these experiments is often much longer and more convergence can
This is probably the first type of set-up that was used in analogue ex- therefore be achieved then in a push-type experiment. Thrusts
periments of brittle thrust wedges (e.g. Cadell, 1890) and is still fre- again first form at the backstop and model results will be sensitive
quently used in analogue and numerical studies (e.g. Burbidge and to the shape of the backstop, its height, its deformability, and the
Braun, 2002; Costa and Vendeville, 2002; Cruz et al., 2010; Koyi, presence of an exit slot, similar to the situation for the push-type ex-
1995; Marques and Cobbold, 2002; Marshak and Wilkerson, 1992; periments. It is important to realise that the difference between a
Miyakawa et al., 2010; Nilforoushan and Koyi, 2007; Schreurs et al., push and a pull-type of set-up is in principle only a question of

Thrust wedge set-ups Model examples


a) Push experiment
2 km
v

b) Conveyor belt experiment

c) S-point experiment
Retro-side Pro-side
5 cm

S v

Fig. 6. Examples of set-ups used in numerical and analogue models of thrust wedges. Schematic set-ups in the left column, examples of model studies in the right column. a) In a
push-type experiment, a mobile wall is moved into the model domain. The example from Miyakawa et al. (2010) shows a distinct-element model result after 22.5 km shortening of
a 1 km thick sand-like material over a low friction base. b) In a conveyor belt experiment, material is pulled towards the backstop. The example from Simpson (2011) shows a finite-
element model after 35% convergence of a strain-weakening elastic-plastic material that is initially 5 km thick. c) In an S-point experiment, material is pulled towards a basal dis-
continuity (the S-point). The example from McClay et al. (2004) shows 51 cm shortening of a quartz sand pack initially 2.5 cm thick. All figures in the right column are reproduced
with permission from Elsevier.
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 9

a) Set-up of the shortening experiment Mobile


10 cm vertical wall

10o
sand

1.0 cm sand
0.5 cm sand
1.0 cm sand
0.5 cm microbeads
0.5 cm sand Velocity
discontinuity

b) Numerical results c) Analogue results


Univ. Bern

I2ELVIS

Univ. Parma
LAPEX-2D

Univ. Pavia
Microfem

PFC2D by Kyoto IFP Rueil-Malmaison

Univ. Toronto

Sopale

-40 -30 -20 cm -40 -30 -20 cm

Fig. 7. Results of the GeoMod2004 thrust wedge comparison experiment (Buiter et al., 2006; Schreurs et al., 2006). a) A 3.5 cm high sand package with an embedded microbeads
layer is shortened by a mobile wall moving into the model domain (push-type experiment, Fig. 6a). A pre-existing wedge of sand with a surface slope of 10° lies on top of the sand
layers, next to the mobile wall. b) Numerical results after 14 cm of shortening. I2ELVIS is a finite-difference model, LAPEX-2D a FLAC-type finite-difference model, Microfem and
SOPALE arbitrary Lagrangian Eulerian finite-element models, and PFC2D a distinct-element model. A description of the codes is in Buiter et al. (2006). c) Analogue results after
14 cm of shortening. The sections of Bern and IFP are CT scans through the centre of the model domain, whereas the sections of Parma, Pavia and Toronto are sidewall observations.
More details of the analogue models can be found in Schreurs et al. (2006). Figures slightly modified from Buiter et al. (2006) (with permission for reproduction of the original
figure from the Geological Society of London).

reference frame in 2-D, but that boundary conditions at the lateral for their analogue experiments of wedge formation in quartz sand
walls, parallel to the shortening direction, play a role in 3-D. If the lat- in a model domain of 37 × 28 × 2.6 cm. The distance of influence of
eral boundaries are held fixed, the surface strike of thrusts in a push the lateral walls will depend on the model material, the sidewall fric-
experiment will be convex with respect to the backstop (mobile tion, and the model set-up.
wall), but concave for conveyor experiments (Cubas et al., 2010; S-point experiments were introduced to mimic a convergent mar-
Schreurs et al., 2006). This is caused by rotation of stresses at the lat- gin setting and, as a result, models often consider the entire crust
eral walls, which exert a drag on the model material. 3-D push exper- with part of the lithospheric mantle (Beaumont et al., 1999; Selzer
iments will as a result have steeper surface slopes at the lateral walls, et al., 2008; Willett et al., 1993). This then requires that
whereas conveyor experiments have shallower surface slopes at the temperature-dependent viscous creep is considered for the deeper
lateral walls, compared to the interior. The influence of the lateral parts of the model. However, S-point experiments have also been
walls reduces towards the interior of the experiments. Cubas et al. used on a shallower scale for mainly brittle models (e.g., Bigi et al.,
(2010) find that lateral effects disappear at ca. 8 cm from the walls, 2010; Hoth et al., 2006; McClay et al., 2004; Naylor et al., 2005;
10 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

Simpson, 2010b; Storti et al., 2000) (Fig. 6c). In these experiments, sands, or with different numerical techniques (Fig. 7). The results of
material is dragged towards a discontinuity (the S-point). The set- the compressional wedge experiment show variability in the number
up has as advantage that deformation occurs away from the end of shear zones that develop, shear zone dip angle, and shear zone
walls, though the location of the first thrusts will now be controlled spacing. These ‘observables’ were measured in the same manner, by
by the S-point. In addition, the set-up can be modified, both in nu- two of the authors, and then averaged. The values of the dip angles
merical and analogue models, to include the effects of isostasy differed on average 2° between the measurers, with a maximum dif-
(Hoth et al., 2006; Simpson, 2010b) and subduction of material out ference of 5–6°. Cubas et al. (2010) report a typical difference of 3° for
of the model domain (Beaumont et al., 1999). I would like to empha- repeated dip angle measurements by the same or different mea-
sise that none of the set-up types discussed here is necessarily better surers. In the comparison experiment of Fig. 7, variability was also ob-
than another as this entirely depends on the research hypothesis that served in deformation style, something which is confirmed in more
is to be tested. As always, it is important to keep experimental limita- recent comparison experiments which followed more strictly defined
tions in geometry and boundary conditions in mind when designing set-ups and used identical materials (Buiter et al., 2010; Schreurs
an analogue or numerical experiment. et al., 2010). This variability in style makes it more difficult to com-
pare experiments directly, as thrusts may initiate in different places
4.2. Quantification of model results and influence shear zone dip angle and spacing. The variability also
raises the question how the style of an experiment could be captured
The discussion in the previous sections points to several model re- in a statistically meaningful manner that would enable quantitative
sults that may be compared to geological and geophysical observa- comparisons.
tions and to other models: surface slope and/or taper value, thrust
dip angle, thrust spacing, wedge width, and deformation style. The
latter is a somewhat vague description that includes characteristics 4.3. Examples of wedge model sensitivities
such as toe accretion versus basal accretion, out-of-sequence thrust
development, the development of long versus short thrust sheets, 4.3.1. The role of basal dip
and whether deformation is accommodated by a few or by multiple Already in one of the early wedge mechanics papers, Davis et al.
thrusts. So far there seems to be no general agreement on how to (1983) showed that critical analogue sand models follow the theoret-
measure thrust wedge surface slopes. The surface slope of wedges ical prediction that an increase in basal dip angle will be accompanied
can be measured as an envelope (e.g., Davis et al., 1983), or from a by a decrease in surface dip angle. This is true for low values of surface
line through the tops of thrust sheets (e.g., Buiter et al., 2006; and basal slope (the lower of the two critical taper branches, see also
Nieuwland et al., 2000; Schreurs et al., 2006), whereas Stockmal Fig. 2), for which an increase in basal dip is accompanied by a de-
et al. (2007) show that for their finite-element models a best fit to crease in surface slope and an increase in taper angle. The dip angle
the critical taper solution is obtained by drawing a line through the of potential forward and backward thrusts relative to the base re-
valleys between the thrust breaks at the surface. Measuring the sur- mains the same (Eqs. 3 and 10), but relative to a horizontal reference
face slope as a line through the tops of thrust sheets, Buiter et al. surface, forward sliplines become steeper, while the potential back-
(2006) report an average difference of 3° for measurements by two ward thrusts become shallower. Referenced to a horizontal surface,
measurers. Thrusts in wedge models can be listric in shape with an the difference in dip between forward and backward thrusts there-
increasing dip angle upwards. This is in contrast with the constant fore reduces. This may influence thrust activation and deformation
dip angle that is predicted in the exact solution of critical taper theory style, though few studies have systematically investigated thrust be-
(Dahlen, 1984). The latter theory relaxes the constraint of σ1 parallel haviour above a varying basal incline. Smit et al. (2003) found a
to the wedge surface. In models that have σ1 parallel to the surface change in deformation style for brittle sand wedges that build above
and that have moderate to high basal friction, listric faults are to be basal dips between 0.75° and 3.0°. A shallow base created a wide
expected (Section 2) (Dahlen et al., 1984; Davis et al., 1983; Hafner, wedge with mainly forward thrusts, while a slightly steeper base
1951; Hubbert, 1951). In addition, theoretical dip angles (Coulomb, led to a shorter wedge dominated by forward and backward thrusts.
Arthur, or Roscoe, Eqs. 11–13) are relative to the direction of the max- In their set-up, initial model thickness also increased with increasing
imum compressive stress (σ1), which is not necessarily horizontal basal dip and it can therefore not be determined if the change in
and this should be taken into account when comparing model thrust thrusting style would not partly be related to the change in model
dip angles to theoretical predictions. Shear zone spacing has been thickness. A thicker model may lead to larger thrust spacing
found to increase with model thickness (Gutscher et al., 1998b; (Marshak and Wilkerson, 1992; Panian and Wiltschko, 2007), though
Marshak and Wilkerson, 1992). This could in part be explained by this would not explain the increased tendency for backthrusting with
thrusts developing through reflection of shear bands between the increasing basal dip.
surface and the base of the model, as proposed by Panian and Boyer (1995) drew attention to the wedge building phase before
Wiltschko (2007). In this case, thrust spacing will also depend on criticality is reached. Wedges that need to grow to reach a critical
thrust dip angle. The strength of the material at the base of the taper will tend to be less wide than wedges that are critical or stable
model also exerts a control on thrust spacing, as a strong base leads in an earlier phase. If wedges are growing above a pre-existing in-
to short wedges with closely spaced thrusts (Mulugeta, 1988). For clined base, shallow basal dips could be associated with short
numerical models, resolution also plays a role and continuum models wedges as more shortening is taken up by internal deformation,
with finer resolution show in general more thrusts and smaller thrust whereas steeper dips lead to long wedges. In nature, the dip angle
spacing than models with coarser resolution (Buiter et al., 2006). In of the wedge base can be expected to adjust isostatically to the in-
addition to the above results that can be determined for analogue creasing overburden during wedge building. Isostatic compensation
and numerical models, numerical models can visualise the evolution has been included in numerical and analogue models (Beaumont et
of strain-rate, strain, stress and pressure fields inside the models, al., 1999; Hoth et al., 2006; Simpson, 2010b; Stockmal et al., 2007).
compare values of gravitational rate of work and the internal rate of For these types of models, the critical state will vary during model
dissipation of energy between models, and obtain flow fields and evolution. Stockmal et al. (2007) find that longer thrust sheets de-
pressure–temperature–time paths. velop in their finite-element model with an isostatic base relative
The comparison experiments of Schreurs et al. (2006) and Buiter to an identical model without isostasy. These models also point in
et al. (2006) illustrate the variability that can result when the same the direction of a sensitivity of wedge deformation style to changes
set-up is modelled with different laboratory apparatus and different in basal dip angle.
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 11

4.3.2. The role of basal strength curves is not notably different for these cohesive results in comparison
Natural fold-and-thrust belts are generally characterised by a with the cohesionless results. This may be because differences in sur-
weak basal layer (Chapple, 1978), but the character of this weak face slope for cohesive materials relative to cohesionless materials
base could vary between a shear zone or basal detachment (such as could be of similar magnitude as errors in surface slope measurements
the subduction zone fault underlying accretionary wedges), a layer or variations caused by non-stable surface slopes. The distinct-
with high fluid pressure, or a weak lithological layer (such as salt in element models of Burbidge and Braun (2002) show the largest dis-
the Zagros Mountains). The thickness, strength, and rheology (fric- crepancy between model results and analytical predictions. As
tional or viscous) of the weak base influence the deformation style Nilforoushan et al. (2008) also point out for their sand models, this
of the overlying rocks. Models with discrete or thin basal detach- may be caused by changing properties of the particle assembly (e.g.,
ments have mainly been used to investigate the response of brittle compaction) during model evolution or by a situation in which surface
wedges to changes in frictional strength of the base. This type of slope has not stabilised yet and more convergence would be needed to
models can be compared with critical taper theory, as the theory reach a critical taper.
approximates the weak base of thrust wedges by a discrete frictional The style of deformation also changes with basal friction as
interface. I will discuss these below. Wedge models with a viscous backward thrusts have been shown to be more abundant for a
basal layer often have a low basal strength and many aspects of weak base in comparison with a strong base (Huiqi et al., 1992;
their deformation style can be understood from a simple comparison Lallemand et al., 1994; Mulugeta, 1988). Also, thrust planes are
to wedges building above a weak discrete base. However, in models expected to become increasingly listric with increasing basal friction
that build above a viscous layer, the velocity-dependence of the (Section 2), but this seems not to have been investigated systemati-
viscous stresses and deformation of the viscous detachment itself cally. Burbidge and Braun (2002) found that basal friction in addition
may also influence the wedge model (Bonini, 2007; Costa and influences the style of accretion, with frontal accretion occurring for
Vendeville, 2002; Smit et al., 2003). In addition, most models assume low basal friction values and underthrusting for higher basal friction
a rather abrupt change in strength between the wedge and its base, as values (see also Gutscher et al., 1998b). In their models, underthrust-
an analogue to a transition from natural wedge rocks to a basal de- ing mainly seems to occur for wedges that have ϕb > ϕ and which are
tachment or a weak lithological base. More gradual transitions in therefore outside the existence field of the analytical theory of critical
strength (lithological or rheological in character) seem mainly con- wedges. I would speculate that the underthrusting regime occurs
fined to larger-scale (crust- to lithosphere-scale) models (e.g., when models build an internal basal thrust that would reduce the
Beaumont et al., 1999; Selzer et al., 2008; Willett, 1999b). basal friction experienced by the wedge above it to ϕb = ϕ. However,
A decrease in the value of basal friction in critical wedges leads to a because a weak basal detachment is one of the assumptions of critical
decrease in taper angle (which implies a shallower wedge for models taper theory, this case can only be compared to the theory once the
with a horizontal base), steeper forward thrusts, and shallower back- new basal thrust is weakened by, for example, strain-softening.
ward thrusts (Section 2). Wedges with a weak base can therefore Basal friction in models of accretionary wedges is an approx-
expected to be wide, with a more-or-less symmetric style of thrusting imation of the subduction interface behaviour which cycles through
(Davis and Engelder, 1985; Huiqi et al., 1992), whereas wedges with a stick-and-slip events (Brace and Byerlee, 1966). Wang and Hu
strong base are narrow and dominated by forward thrusts (Fig. 8). (2006) speculate that the topography of the outer, most seaward,
Several analogue and numerical models confirm that a decrease in wedge is set during large earthquakes, when friction on this
basal strength decreases the critical taper (Burbidge and Braun, velocity-strengthening portion of the subduction fault is high. In
2002; Gutscher et al., 1998b; Koyi and Vendeville, 2003; Lallemand that case, topography and taper angle might not be directly represen-
et al., 1994; Mulugeta, 1988; Naylor et al., 2005; Nieuwland et al., tative of a time-averaged basal friction value, but instead relate to the
2000; Nilforoushan et al., 2008; Teixell and Koyi, 2003; Willett, maximum basal friction value experienced on the underlying subduc-
1992). Quantitatively, several of these model results can be shown to tion interface.
fit the predictions of critical taper theory (Fig. 9). The analytical curves In summary, the effect of variations in strength of the basal de-
in Fig. 9 are for a dry, cohesionless wedge, whereas 7 of the 20 model tachment seems to be relatively well understood. A weak base leads
results are for a cohesive material, with cohesion varying between 20 to a wedge built through frontal accretion with a shallow taper
and 200 Pa (cm-scale). The discrepancy or agreement to the analytical angle and a more symmetric deformation style in which the

a) High basal friction

b) Intermediate basal friction

5 cm

c) Low basal friction

Fig. 8. Example of the effect of basal friction on a sand wedge, redrawn from Huiqi et al. (1992). a) Basal friction 29°, basal cohesion 71 Pa, 66% contraction, b) basal friction 25°, basal
cohesion 25 Pa, 44% contraction c) basal friction 20°, basal cohesion 19 Pa, 26.5% contraction. The sand has an internal angle of friction of 29° and a cohesion of 190 Pa.
12 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

0o 10o 20o 30o 40o


18o
b) B27
o B39
o

o
M19

35 o

40 o

45 o
25 o

30 o
o
14

15o

20 o
o
B11
surface slope α

o
N17
o
M19 M25
o

10o
o
Bi29
o
B39 o o
18o
Ni30 K30
a) 20o 25o 30o 35o 40o 45o
o
B39 o o
N17 H29
o
14 φ = 15 o

surface slope α
6o
o
10o
N17
o
o K30
S12 o
Si30 B27 6o
o

2o B27
o
β = 0o
o λ=0
2
C=0
0o 10o 20o 30o 0o 10o 20o 30o 40o
φb φb

Fig. 9. a) Surface slope versus basal friction for a cohesionless dry critical wedge with a horizontal base. The curves are labelled for increasing values (from left to right) of angle of
internal friction. The lowest of the two permissible critical taper angles is plotted (Fig. 2). b) Numerical and analogue model results plotted on the diagram of a). Letters denote the
study: B = distinct-element models of Burbidge and Braun (2002) with C = 0 Pa, Bi = analogue model of Bigi et al. (2010) with C = 0 Pa, H = analogue model of Huiqi et al. (1992)
with C = 190 Pa (ϕ softening from 29° to 27°), K = analogue models of Konstantinovskaia and Malavieille (2005) with C = 20 Pa, M = analogue models of Mulugeta (1988) with
C = 0 Pa except C = 176 Pa for M25°, N = distinct-element models of Naylor et al. (2005) (with unknown cohesion), Ni = analogue model of Nilforoushan et al. (2008) with
C = 140 Pa, S = finite-element model of Selzer et al. (2007) with C = 10 MPa (model is on km-scale and C is approximately 100 Pa on cm-scale; ϕ softening from 12° to 6°), and
Si = finite-element model of Simpson (2011) with C = 20 MPa strain-weakening to 0 MPa (model on km-scale, C is approximately 200 Pa on cm-scale). The number after each letter
gives the angle of internal friction ϕ of the model experiment, which is constant unless otherwise noted. All surface slope measurements taken from the respective studies, except
for Burbidge and Braun (2002), where I measured surface slope from a line connecting the topographic lows.

difference in dip angle between forward and backward thrusts is re- overburden. Erosion and sedimentation may both lower the wedge
duced. In contrast, a strong base may cause underthrusting and taper angle and bring a wedge into the unstable field. The wedge
leads to a steep wedge with shallow forward thrusts and steep back- will then need to deform, possibly through out-of-sequence thrust-
ward thrusts. ing, to return to its critical taper angle.
The response of brittle thrust wedges to various types of surface
4.3.3. The role of surface processes processes has been investigated in several analogue and numerical
Erosion and sedimentation of surface materials directly change to- model studies that have resulted in a fairly good understanding of
pography and have been shown to have a feedback effect on internal the above controlling factors (Fig. 10). Erosion and sedimentation
deformation (Beaumont et al., 1992; Hilley and Strecker, 2004; processes are necessarily simplified in models and perhaps most so
Koons, 1990; Willett, 1999a). This feedback occurs through different in analogue studies that simulate discrete erosion and sedimentation
mechanisms. Firstly, surface processes affect gravitational stresses events. However, as pointed out by Hoth et al. (2006), also in nature
through a change of the overburden. For brittle wedges, this can erosion may often be non-continuous and episodic (for example,
lead to an increase in basal stresses for sedimentation (and a decrease through landslides and floods). Analogue wedge models have
for the case of erosion). Changes in the gravitational stresses also in- employed inclined (Konstantinovskaia and Malavieille, 2005) or hor-
fluence brittle strength through the associated change in dynamic izontal (Bigi et al., 2010; Persson and Sokoutis, 2002) erosion planes,
pressure. Erosion reduces dynamic pressure and can be seen as a erosion rate linearly increasing with elevation (Hoth et al., 2006) or
weakening process, whereas sedimentation is strengthening. Second- following a non-linear fluvial bedrock incision rule (Cruz et al.,
ly, surface processes can change the taper angle of brittle wedges, 2010). Numerical wedge models have used slope-dependent erosion
through the addition or removal of material at the surface and models (Stockmal et al., 2007), erosion rate depending on slope and
through an isostatic response of the base to changes in the discharge (Willett, 1999a), diffusive erosion and sedimentation

a) Effect of sedimentation (Storti and McClay, 1995) b) Effect of erosion (Cruz et al., 2010)
1) No sedimentation 1) No erosion

5 cm 1 cm

2) Intermediate sedimentation
2) High erosion

Fig. 10. a) Example of the effects of syntectonic sedimentation on a thrust wedge model. Alternate layers of quartz sand (ϕ = 29°) and mica flakes (ϕ = 23°) are shortened against a
backstop in a conveyor-belt set-up (Fig. 6b). Basal friction is 25°. Sand sediments (light grey) were added every centrimetre of contraction. Top model without sedimentation (at
57% contraction), below model with intermediate sedimentation (at 32% contraction). Redrawn from Storti and McClay (1995). Yellow layer is a marker layer for visualisation
added in the redrawing. b) Example of syntectonic erosion on a thrust wedge in a mobile backstop experiment (Fig. 6a). The model consists of 28 mm of quartz sand (ϕ = 33°)
overlying 2 mm of glass beads (ϕb = 25°). Erosion is applied every 7 cm of displacement by following a bedrock incision rule (Hilley et al., 2004). Model with negligible erosion
at top and model with high erosion intensity at the bottom. Both models at 50 cm displacement. Redrawn from Cruz et al. (2010).
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 13

(Selzer et al., 2007; Simpson, 2006), and sedimentation that fills or addition, pre-existing heterogeneities may localise deformation and in-
overfills basins (Simpson, 2010a; Stockmal et al., 2007). The appro- fluence thrusting style (Sassi et al., 1993; Selzer et al., 2007).
priate choice of surface process law is not trivial as the location and Fluids have the effect of lowering effective stress, which promotes the
intensity of erosion and sedimentation exert an important control onset of failure. Wet wedges are therefore expected to be weaker, with a
on wedge deformation style (Cruz et al., 2008; Hoth et al., 2006; higher taper angle. Strayer et al. (2001) coupled a fluid flow model to a
Simpson, 2006; Storti and McClay, 1995). compressional finite-element model of thrust wedge formation. They
Brittle thrust wedge models have shown that erosion maintains find that the presence of water facilitates failure in the wedge, reduces
shear zone activity, localises deformation in the eroding domain basal friction and, focusses fluids within fault zones, which is in agree-
(Bigi et al., 2010; Konstantinovskaia and Malavieille, 2005; ment with analytical predictions. Cobbold et al. (2001) and Mourgues
Konstantinovskaya and Malavieille, 2011; Selzer et al., 2007; Willett, and Cobbold (2006) pioneered the inclusion of fluids in analogue sand
1999a), reduces wedge width (Cruz et al., 2010; Dahlen and Suppe, models, using compressed air as a pore fluid. Their results show that de-
1988), and focuses particle trajectories and exhumation (Fig. 10b). tachments form beneath sealing layers of low permeability, where large
Doubly-vergent wedges that develop over velocity discontinuities il- fluid pressures reduce brittle strength. Décollement levels were found to
lustrate how exhumation patterns focus towards the pro-side for be of less importance in the S-point sand models of Hoth et al. (2007),
pro-wedge erosion and towards the retro-side for retro-wedge ero- but drastically influence deformation style in the analogue and numerical
sion (see Fig. 6c for a definition of pro- and retro-sides) (Cruz et al., models of, among others, Kukowski et al. (2002); Selzer et al. (2007);
2008; Hoth et al., 2006; Willett, 1999a; Willett et al., 1993). Sedimen- Stockmal et al. (2007) and Konstantinovskaya and Malavieille (2011).
tation in turn has a delocalising effect on deformation through Yamato et al. (2011) show how several internal weak layers, represent-
strengthening of brittle material under increasing overburden ing evaporites or shales, within a brittle layer bring their model to a
(Selzer et al., 2007). In contrast, analogue mobile backstop models folding-dominated mode, as opposed to thrust-dominated for brittle
have found that sedimentation focuses deformation towards the models which have a weak layer only at their base. As a general rule, a
backstop (Storti and McClay, 1995). This is not necessarily in dis- décollement needs to be weaker than the base in order to be activated.
agreement with the expected delocalisation, as the area with shear Other factors that may influence the activation of décollements are
zones near the backstop may be the weaker region relative to the ad- their strength difference relative to the surrounding materials, their
jacent undeformed domain that is covered with sediments (Fig. 10a). thickness, and their orientation with respect to the stress field. Models
This again illustrates that wedge evolution will not only be sensitive with décollements can show alternate activation between the internal
to the amount of erosion or sedimentation, but also to where erosion weak layer and the base, accompanied by basal and frontal accretion,
or sedimentation occurs. flat-ramp-flat shear zone geometries, duplexes, and anti-formal stacks
(Fig. 11). These structures cannot be captured by critical taper theory,
4.3.4. The role of internal strength variations which describes an overall, average behaviour of fold-and-thrust belts,
An increase in bulk strength increases the stability field for critical nor by analogue and numerical models of homogeneous, isotropic mate-
wedges, thereby reducing the lower critical surface slope and increasing rials. Results of inhomogeneous dynamic models illustrate how some of
the higher critical surface slope value (Fig. 2f). As most wedges have the complex internal behaviour that is observed in natural fold-and-
small tapers following the lower critical taper solution, this means that thrust belts can be captured and investigated with numerical and ana-
wedges can support a larger taper angle as they become weaker. This logue models.
may perhaps be counterintuitive, but can be understood by considering
a decrease in internal strength as a relative increase in basal strength 5. Concluding remarks
(Section 4.3.2).
Internal strength variations in fold-and-thrust belts and accre- A review of brittle thrust wedge models can hardly do justice to
tionary wedges may occur because of variations in lithology (e.g., the rich literature that this topic has generated. Scaled analogue ex-
evaporite or shale layers), rheology (brittle or viscous), degree of periments have been used to investigate brittle compressional tec-
compaction, lithification, and fluid pressures. Material compacts as tonics for over a century (Cadell, 1890; Willis, 1893) and are still
it is further incorporated into a wedge, which causes an increase in actively used as testified by recent research papers (e.g., Bigi et al.,
wedge strength through an increase in cohesion. Zhao et al. (1986) 2010; Cruz et al., 2010; Cubas et al., 2010). It may in fact seem that an-
show that this would be accompanied by a reduction in critical alogue models are a preferred tool, as the number of numerical stud-
taper angle and, therefore, the creation of a wedge with a convex ies of brittle thrust wedges is lagging behind the number of analogue
shape. In their application to the Barbados wedge, Zhao et al. (1986) studies. This may perhaps be for a historical reason as analogue
derive a lateral increase in cohesion from negligible at the wedge models build on a long tradition of 3-D scaled experiments, whereas
toe to ca. 15 MPa at 50–70 km away from the toe, where wedge thick- numerical models that can accurately resolve shear zones along
ness is about 3 km. The role of cohesion will depend on the magni- which deformation can be accommodated have only become avail-
tude of the frictional strength, the reduction of frictional strength by able in the last few decades (see examples in Buiter et al., 2006;
pore fluid pressures, and the depth interval under consideration. For Ellis et al., 2004; Miyakawa et al., 2010; Simpson, 2011; Stockmal et
example, using dry elastic–plastic finite-element models, Simpson al., 2007). Numerical and analogue models could be combined to mu-
(2011) finds only a minimal influence of cohesion on deformation tual benefit, merging the ease of 3-D experiments in the laboratory
on the scale of the entire fold-and-thrust belt. with accurate sensitivity analyses in numerical models.
Compaction and dilation also affect the frictional strength of loose Many aspects of fold-and-thrust belts and accretionary wedges
granular materials (Fig. 4). Lohrmann et al. (2003) infer that the bulk can be understood from the analytical critical taper theory, which
strength of critical wedge segments is controlled by the frictional considers wedge formation in a homogeneous material that is on
strength of its weakest part, which are the faults that have reached the verge of brittle failure throughout. Analogue and numerical stud-
their stable sliding strength. This then implies that comparisons of ies of thrust wedges have confirmed predictions from critical taper
model results to critical taper theory should really be made using the sta- theory that (see Section 4.3 and references therein):
ble sliding value for angle of internal friction instead of the peak value
which is often used. As the properties of analogue brittle materials de- • an increase in basal dip angle causes a decrease in surface dip angle,
pend on material type and handling, a certain degree of variability is to • a decrease in basal strength leads to a shallower wedge and a more
be expected for models that use similar set-ups, but different materials symmetric thrust style through steeper forward thrusts, and shal-
(Lohrmann et al., 2003; Mulugeta, 1988; Schreurs et al., 2006). In lower backward thrusts,
14 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

a) Basal detachment layer

20 km

b) Basal and internal detachment layers


detachments

Fig. 11. Example of two finite-element models from Stockmal et al. (2007) that illustrate the influence of an internal detachment layer. A 5 km thick viscous-plastic layer (ϕ = 38°,
C = 2 MPa) is dragged over a weak base (ϕ = 3.5°, C = 2 MPa) towards a rigid backstop to the left. Basal flexure develops in response to the increasing overburden. a) Model with a
basal detachment layer. b) Model with a basal detachment in combination with an internal detachment (ϕ = 1°, C = 2 MPa). The internal detachment layer favours the formation of
longer thrust sheets. Convergence is 240 km in both models. Figures reproduced from Stockmal et al. (2007) with permission from the Geological Society of America.

• erosion promotes narrower thrust belts and exhumation, and The steadily improving resemblance to internal structures of nat-
• an increase in bulk strength decreases surface dip angle ural thrust wedges that is achieved by increasingly complex numeri-
cal and analogue studies raises the question whether a ‘minimum set’
These trends are, therefore, robust results of wedge studies that of requirements could be found that would capture the average
can confidently be used in interpretations of natural processes. A longer-term behaviour of wedges (as in critical taper theory) simulta-
comparison of the wedge studies discussed in this review indicates neously with bridging the scale between detailed upper crustal struc-
that thrust vergence, thrust sheet length, surface slope, basal dip tures and larger-scale lithosphere and mantle behaviour. This would
angle, and wedge width could be classified as fairly reliable model then perhaps allow us to take the field a step further and use observa-
outcomes, whereas detailed thrust dip angle, number of thrusts, and tions of natural wedges to constrain the deformational processes that
thrust spacing are variable between models, depending not only on are involved in creating thrust belts and accretionary wedges.
the physics of the problem, but also on the analogue material and nu-
merical resolution (see also Buiter et al., 2006; Schreurs et al., 2006).
Acknowledgements
As critical taper theory describes an average, long-term behaviour
of thrust wedges, analogue and numerical models can be used to
I gratefully acknowledge funding by the Norwegian Research
take the field a step further by analysing transient behaviour, non-
Council (NFR project 180449). I would like to thank Fabrizio Storti
homogeneous materials, and internal deformation styles. Modelling
and Timothy Horscroft for the manuscript invitation, Guido Schreurs
studies have, for example, (a) investigated wedges beyond the
for many discussions on brittle wedges and analogue modelling tech-
existence field of critical taper solutions (Burbidge and Braun,
niques, and Steffen Abe, David Egholm, Marc-André Gutscher, Boris
2002), (b) illustrated the role of internal detachment layers and ma-
Kaus, Vincent Lemiale, Louis Moresi, and Tim Redfield for discussions
terial heterogeneities (Konstantinovskaya and Malavieille, 2011;
on various aspects of brittle behaviour. Matthias Klinkmüller send a
Kukowski et al., 2002; Selzer et al., 2007; Stockmal et al., 2007), (c)
preprint of his manuscript on analogue material properties, and
shown that surface erosion localises deformation, maintains shear
Ayumu Miyakawa and Guy Simpson provided images of their pub-
zone activity, and promotes steep shear zones (e.g., Bigi et al., 2010;
lished model results reproduced in Fig. 6. Many thanks to Matthias
Cruz et al., 2010; Konstantinovskaia and Malavieille, 2005; Willett,
Rosenau and Boris Kaus for their constructive reviews.
1999a), (d) found changes in thrust style with basal dip changes
(Smit et al., 2003; Stockmal et al., 2007), and (e) shown that a certain
amount of convergence may be needed (which may perhaps vary per References
method) to achieve a critical taper. Abe, S., Mair, K., 2009. Effects of gouge fragment shape on fault friction: new 3D model-
Some open questions that seem to have been little addressed so ling results. Geophysical Research Letters 36, L23302. doi:10.1029/2009GL040684.
far in wedge dynamics studies concern: Adam, J., Urai, J.L., Wieneke, B., Oncken, O., Pfeiffer, K., Kukowski, N., Lohrmann, J., Hoth,
S., van der Zee, W., Schmatz, J., 2005. Shear localisation and strain distribution dur-
ing tectonic faulting—new insights from granular-flow experiments and high-
• the role of basal cohesion (if any) in wedge deformation,
resolution optical image correlation techniques. Journal of Structural Geology 27,
• whether thrusts are indeed more listric for higher basal friction, 283–301. doi:10.1016/j.jsg.2004.08.008.
• whether there is a dependence of shear zone dip angles in distinct- Agarwal, K.K., Agrawal, G.K., 2002. Analogue sandbox models of thrust wedges with
element models on particle size, variable basal frictions. Gondwana Research 5, 641–647.
Antonellini, M.A., Pollard, D.D., 1995. Distinct element modeling of deformation bands
• if different numerical techniques for implementing basal and side in sandstone. Journal of Structural Geology 17, 1165–1182.
friction result in the same behaviour (i.e., a benchmark study), Arthur, J.R.F., Dunstan, T., Al-Ani, Q.A.J., Assadi, A., 1977. Plastic deformation and failure
• which factors and processes control thrust sheet length and the role of granular media. Géotechnique 27, 53–74.
Bally, A.W., Gordy, P.L., Stewart, G.A., 1966. Structure, seismic data, and orogenic evo-
of isostasy therein, and lution of Southern Canadian Rocky Mountains. Bulletin of Canadian Petroleum Ge-
• the level of model complexity that is required to reproduce struc- ology 14, 337–381.
tures in natural wedges. Barcilon, V., 1987. A note on ‘Noncohesive critical Coulomb wedges: an exact solution’
by F.A. Dahlen. Journal of Geophysical Research 92, 3681–3682.
Bardet, J.P., Proubet, J., 1991. A numerical investigation of the structure of persistent
Many thrust wedges are built on former passive margins, with
shear bands in granular media. Géotechnique 41, 599–613.
pre-existing normal faults, and lateral and vertical variations in lithol- Beaumont, C., Fullsack, P., Hamilton, J., 1992. Erosional control of active compressional
ogies. Does the complexity of structures that are observed in many orogens. In: McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall, New York, pp.
natural fold-and-thrust belts and accretionary wedges require 1–18.
Beaumont, C., Ellis, S., Pfiffner, A., 1999. Dynamics of sediment subduction–accretion at
lithological layering, pre-existing structures, and a wide range of convergent margins: short-term modes, long-term deformation, and tectonic im-
rheologies? plications. Journal of Geophysical Research 104, 17,573–17,601.
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 15

Bigi, S., Di Paolo, L., Vadacca, L., Gambardella, G., 2010. Load and unload as interference Egholm, D.L., 2007. A new strategy for discrete element numerical models: 1. Theory.
factors on cyclical behavior and kinematics of Coulomb wedges: insights from Journal of Geophysical Research 112, B05203. doi:10.1029/2006JB004557.
sandbox experiments. Journal of Structural Geology 32, 28–44. doi:10.1016/ Eisenstadt, G., Sims, D., 2005. Evaluating sand and clay models: do rheological differ-
j.jsg.2009.06.018. ences matter? Journal of Structural Geology 27, 1399–1412. doi:10.1016/
Bonini, M., 2007. Deformation patterns and structural vergence in brittle–ductile thrust j.jsg.2005.04.010.
wedges: an additional analogue modelling perspective. Journal of Structural Geol- Ellis, S., Schreurs, G., Panien, M., 2004. Comparisons between analogue and numerical
ogy 29, 141–158. doi:10.1016/j.jsg.2006.06.012. models of thrust wedge development. Journal of Structural Geology 26, 1659–1675.
Boyer, S.E., 1995. Sedimentary basin taper as a factor controlling the geometry and ad- Emerman, S.H., Turcotte, D.L., 1983. A fluid model for the shape of accretionary wedges.
vance of thrust belts. American Journal of Science 295, 1220–1254. Earth and Planetary Science Letters 63, 379–384.
Brace, W.F., Byerlee, J.D., 1966. Stick-slip as a mechanism for earthquakes. Science 153, Gerya, T., 2010. Dynamical instability produces transform faults at mid-ocean ridges.
990–992. Science 329, 1047–1050. doi:10.1126/science.1191349.
Braun, J., Yamato, P., 2010. Structural evolution of a three-dimensional, finite-width Glodny, J., Lohrmann, J., Echtler, H., Gräfe, K., Seifert, W., Collao, S., Figueroa, O.,
crustal wedge. Tectonophysics 484, 181–192. doi:10.1016/j.tecto.2009.08.032. 2005. Internal dynamics of a paleoaccretionary wedge: insights from combined
Buiter, S.J.H., Babeyko, A.Yu., Ellis, S., Gerya, T.V., Kaus, B.J.P., Kellner, A., Schreurs, G., isotope tectonochronology and sandbox modelling of the South-Central Chilean
Yamada, Y., 2006. The numerical sandbox: comparison of model results for a short- forearc. Earth and Planetary Science Letters 231, 23–39. doi:10.1016/
ening and an extension experiment. In: Buiter, S.J.H., Schreurs, G. (Eds.), Analogue j.epsl.2004.12.014.
and Numerical Modelling of Crustal-Scale Processes: Geol. Soc. London Spec. Publ., Gutscher, M.-A., Kukowski, N., Malavieille, J., Lallemand, S., 1998a. Episodic imbricate
253, pp. 29–64. thrusting and underthrusting: analog experiments and mechanical analysis ap-
Buiter, S., Albertz, M., Cooke, M., Crook, T., Egholm, D., Ellis, S., Gerya, T., Hodkinson, L., plied to the Alaskan accretionary wedge. Journal of Geophysical Research 103,
Kaus, B., Landry, W., Maillot, B., Mishin, Y., Pascal, C., Schreurs, G., Souloumiac, P., 10,161–10,176.
Beaumont, C., 2010. Quantitative comparisons of numerical models of brittle Gutscher, M.-A., Kukowski, N., Malavieille, J., Lallemand, S., 1998b. Material transfer in
wedge dynamics. Geophysical Research Abstracts 12, EGU2010-12325. accretionary wedges from analysis of a systematic series of analog experiments.
Burbidge, D.R., Braun, J., 2002. Numerical models of the evolution of accretionary Journal of Structural Geology 20, 407–416.
wedges and fold-and-thrust belts using the distinct-element method. Geophysical Hafner, W., 1951. Stress distributions and faulting. Bulletin of the Geological Society of
Journal International 148, 542–561. America 62, 373–398.
Butler, R.W.H., 1982. The terminology of structures in thrust belts. Journal of Structural Han, C., Drescher, A., 1993. Shear bands in biaxial tests on dry coarse sands. Soils and
Geology 4, 239–245. Foundations 33, 118–132.
Byerlee, J., 1978. Friction of rocks. Pageoph 116, 615–626. Haq, S.S.B., Davis, D.M., 2008. Extension during active collision in thin-skinned wedges:
Byrne, D.E., Wang, W.-H., Davis, D.M., 1993. Mechanical role of backstops in the growth insights from laboratory experiments. Geology 36, 475–478.
of forearcs. Tectonics 12, 123–144. Hermann, H.J., 2001. Structures in deformed granular packings. Granular Matter 3, 15–18.
Cadell, H.M., 1890. Experimental researches in mountain building. Transactions of the Hilley, G.E., Strecker, M.R., 2004. Steady state erosion of critical Coulomb wedges with
Royal Society of Edinburgh 35, 337–357. applications to Taiwan and the Himalaya. Journal of Geophysical Research 109,
Chapple, W.M., 1978. Mechanics of thin-skinned fold-and-thrust belts. Geological Soci- B01411. doi:10.1029/2002JB002284.
ety of America Bulletin 89, 1189–1198. Hilley, G.E., Strecker, M.R., Ramos, V.A., 2004. Growth and erosion of fold-and-thrust
Cloos, E., 1955. Experimental analysis of fracture patterns. Bulletin of the Geological So- belts with an application to the Aconcagua fold-and-thrust belt, Argentina. Journal
ciety of America 66, 241–256. of Geophysical Research 109, B01410. doi:10.1029/2002JB002282.
Cobbold, P.R., Durand, S., Mourgues, R., 2001. Sandbox modelling of thrust wedges with Hoth, S., Adam, J., Kukowski, N., Oncken, O., 2006. Influence of erosion on the kinemat-
fluid-assisted detachments. Tectonophysics 334, 245–258. ics of bivergent orogens: results from scaled sandbox simulations. Geological Soci-
Colletta, B., Letouzey, J., Pinedo, R., Ballard, J.F., Balé, P., 1991. Computerized X-ray to- ety of America—Special Paper 398, 201–225. doi:10.1130/2006.2398(12).
mography analysis of sandbox models: examples of thin-skinned thrust systems. Hoth, S., Hoffmann-Rothe, A., Kukowski, N., 2007. Frontal accretion: an internal clock
Geology 19, 1063–1067. for bivergent wedge deformation and surface uplift. Journal of Geophysical Re-
Cooper, M., 2007. Structural style and hydrocarbon prospectivity in fold and thrust search 112, B06408. doi:10.1029/2006JB004357.
belts: a global review. In: Ries, A.C., Butler, R.W.H., Graham, R.H. (Eds.), Deforma- Hubbert, M.K., 1937. Theory of scale models as applied to the study of geologic struc-
tion of the Continental Crust: The Legacy of Mike Coward: Geol. Soc., Lond., Spec. tures. Bulletin of the Geological Society of America 48, 1459–1519.
Publ., 272, pp. 447–472. Hubbert, M.K., 1951. Mechanical basis for certain familiar geologic structures. Bulletin
Costa, E., Vendeville, B.C., 2002. Experimental insights on the geometry and kinematics of the Geological Society of America 62, 355–372.
of fold-and-thrust belts above weak, viscous evaporitic décollement. Journal of Hubbert, M.K., Rubey, W.W., 1959. Role of fluid pressure in mechanics of overthrust
Structural Geology 24, 1729–1739. faulting. 1. Mechanics of fluid-filled porous solids and its application to overthrust
Coulomb, C.A., 1773. L'application des règles de maximis et minimis à quelques problè- faulting. Bulletin of the Geological Society of America 70, 115–166.
mes de statique, relatifs à l'architecture. Mémoires de Mathématique et de Phy- Huiqi, Lui, McClay, K.R., Powell, D., 1992. Physical models of thrust wedges. In: McClay,
sique, Academie Royale des Sciences 7, 343–382. K.R. (Ed.), Thrust Tectonics. Chapman and Hall, New York, pp. 71–81.
Coward, M.P., 1983. Thrust tectonics, thin-skinned or thick-skinned and the continua- Jaeger, J.C., Cook, N.G.W., Zimmerman, R.W., 2007. Fundamentals of Rock Mechanics,
tion of thrusts to deep in the crust. Journal of Structural Geology 5, 113–125. fourth edition. Blackwell Publishing. 475 pp.
Cruz, L., Teyssier, C., Perg, L., Take, A., Fayon, A., 2008. Deformation, exhumation, and topog- Kaus, B.J.P., 2010. Factors that control the angle of shear bands in geodynamic numer-
raphy of experimental doubly-vergent orogenic wedges subjected to asymmetric ero- ical models of brittle deformation. Tectonophysics 484, 36–47. doi:10.1016/
sion. Journal of Structural Geology 30, 98–115. doi:10.1016/j.jsg.2007.10.003. j.tecto.2009.08.042.
Cruz, L., Malinski, J., Wilson, A., Take, W.A., Hilley, G., 2010. Erosional control of the kinemat- Klinkmüller, M., 2011. Properties of analogue materials, experimental reproducibility
ics and geometry of fold-and-thrust belts imaged in a physical and numerical sandbox. and 2D/3D deformation quantification techniques in analogue modelling of crust-
Journal of Geophysical Research 115, B09404. doi:10.1029/2010JB007472. al-scale processes. Ph.D. Thesis, University of Bern.
Cubas, N., Maillot, B., Barnes, C., 2010. Statistical analysis of an experimental compres- Konstantinovskaia, E., Malavieille, J., 2005. Erosion and exhumation in accretionary
sional sand wedge. Journal of Structural Geology 32, 818–831. orogens: experimental and geological approaches. Geochemistry, Geophysics,
Cundall, P., Strack, O., 1979. A discrete numerical model for granular assemblies. Géo- Geosystems 6, Q02006. doi:10.1029/ 2004GC000794.
technique 29, 47–65. Konstantinovskaya, E., Malavieille, J., 2011. Thrust wedges with décollement levels and
Cuvelier, C., Segal, A., van Steenhoven, A.A., 1986. Finite Element Methods and Navier– syntectonic erosion: a view from analog models. Tectonophysics 502, 336–350.
Stokes Equations. D. Reidel Publishing Company, Dordrecht, Holland. 483 pp. doi:10.1016/j.tecto.2011.01.020.
Dahlen, F.A., 1984. Noncohesive critical Coulomb wedges: an exact solution. Journal of Koons, P.O., 1990. Two-sided orogen: collision and erosion from the sandbox to the
Geophysical Research 89, 10,125–10,133. Southern Alps, New Zealand. Geology 18, 679–682.
Dahlen, F.A., 1990. Critical taper model of fold-and-thrust belts and accretionary Koyi, H., 1995. Mode of internal deformation in sand wedges. Journal of Structural Ge-
wedges. Annual Review of Earth and Planetary Sciences 18, 55–99. ology 17, 293–300.
Dahlen, F.A., Barr, T., 1989. Brittle frictional mountain building 1. Deformation and me- Koyi, H.A., Mancktelow, N.S. (Eds.), 2001. Tectonic modeling: a volume in honor of
chanical energy budget. Journal of Geophysical Research 94, 3906–3922. Hans Ramberg: Geol. Soc. Am. Mem., 193. 286 pp.
Dahlen, F.A., Suppe, J., 1988. Mechanics, growth, and erosion of mountain belts. Geolog- Koyi, H.A., Vendeville, B.C., 2003. The effect of décollement dip on geometry and kine-
ical Society of America—Special Paper 218, 161–178. matics of model accretionary wedges. Journal of Structural Geology 25, 1445–1450.
Dahlen, F.A., Suppe, J., Davis, D., 1984. Mechanics of fold-and-thrust belts and accre- Krantz, R.W., 1991. Measurements of friction coefficients and cohesion for faulting and
tionary wedges: cohesive Coulomb theory. Journal of Geophysical Research 89, fault reactivation in laboratory models using sand and sand mixtures. Tectonophy-
10,087–10,101. sics 188, 203–207.
Davis, D.M., Engelder, T., 1985. The role of salt in fold-and-thrust belts. Tectonophysics Kukowski, N., Lallemand, S.E., Malavieille, J., Gutscher, M.-A., Reston, T.J., 2002. Me-
119, 67–88. chanical decoupling and basal duplex formation observed in sandbox experiments
Davis, D., Suppe, J., Dahlen, F.A., 1983. Mechanics of fold-and-thrust belts and accre- with application to the Western Mediterranean Ridge accretionary complex. Ma-
tionary wedges. Journal of Geophysical Research 88, 1153–1172. rine Geology 186, 29–42.
de Borst, R., Sluys, L.J., 1991. Localisation in a Cosserat continuum under static and dy- Lallemand, S.E., Schnürle, P., Malavieille, J., 1994. Coulomb theory applied to accretion-
namic loading conditions. Computer Methods in Applied Mechanics and Engineer- ary and nonaccretionary wedges: possible causes for tectonic erosion and/or fron-
ing 90, 805–827. tal accretion. Journal of Geophysical Research 99, 12,033–12,055.
Del Castello, M., McClay, K.R., Pini, G.A., 2005. Role of preexisting topography and over- Lemiale, V., Mühlhaus, H.-B., Moresi, L., Stafford, J., 2008. Shear banding analysis of
burden on strain partitioning of oblique doubly vergent convergent wedges. Tec- plastic models formulated for incompressible viscous flows. Physics of the Earth
tonics 24, TC6004. doi:10.1029/2005TC001816. and Planetary Science Letters 171, 177–186. doi:10.1016/j.pepi.2008.07.038.
16 S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17

Lohrmann, J., Kukowski, N., Adam, J., Oncken, O., 2003. The impact of analogue material Rodgers, J., 1949. Evolution of thought on structure of middle and southern Appala-
properties on the geometry, kinematics, and dynamics of convergent sand wedges. chians. American Association of Petroleum Geologists Bulletin 33, 1643–1654.
Journal of Structural Geology 25, 1691–1711. Roscoe, K.H., 1970. The influence of strains in soil mechanics. Géotechnique 20, 129–170.
Luján, M., Rossetti, F., Storti, F., Ranalli, G., Socquet, A., 2010. Flow trajectories in ana- Rossetti, F., Faccenna, C., Ranalli, G., Storti, F., 2000. Convergence rate-dependent
logue viscous orogenic wedges: insights on natural orogens. Tectonophysics 484, growth of experimental viscous orogenic wedges. Earth and Planetary Science Let-
119–126. doi:10.1016/ j.tecto.2009.09.009. ters 178, 367–372.
Mair, K., Frye, K.M., Marone, C., 2002. Influence of grain characteristics on the friction of Rossi, D., Storti, F., 2003. New artificial granular materials for analogue laboratory ex-
granular shear zones. Journal of Geophysical Research 107, 2219. doi:10.1029/ periments: aluminium and siliceous microspheres. Journal of Structural Geology
2001JB000516. 25, 1893–1899. doi:10.1016/S0191-8141(03)00041-5.
Marques, F.O., Cobbold, P.R., 2002. Topography as a major factor in the development of arcu- Sassi, W., Colletta, B., Balé, P., Paquereau, T., 1993. Modelling of structural complexity in
ate thrust belts: insights from sandbox experiments. Tectonophysics 348, 247–268. sedimentary basins: the role of pre-existing faults in thrust tectonics. Tectonophy-
Marshak, S., Wilkerson, M.S., 1992. Effect of overburden thickness on thrust belt geom- sics 226, 97–112.
etry and development. Tectonics 11, 560–566. Schanz, T., Vermeer, P.A., 1996. Angles of friction and dilatancy of sand. Géotechnique
McClay, K.R., Whitehouse, P.S., Dooley, T., Richards, M., 2004. 3D evolution of fold and 46, 145–151.
thrust belts formed by oblique convergence. Marine and Petroleum Geology 21, Schellart, W.P., 2000. Shear test results for cohesion and friction coefficients for differ-
857–877. ent granular materials: scaling implications for their usage in analogue modelling.
Medvedev, S., 2002. Mechanics of viscous wedges: modeling by analytical and numer- Tectonophysics 324, 1–16.
ical approaches. Journal of Geophysical Research 107. doi:10.1029/2001JB000145. Schreurs, G., Hänni, R., Vock, P., 2001. Four-dimensional analysis of analog models: ex-
Melosh, H.J., Williams, C.A., 1989. Mechanics of graben formation in crustal rocks: a fi- periments on transfer zones in fold and thrust belts. In: Koyi, H.A., Mancktelow,
nite element analysis. Journal of Geophysical Research 94, 13,961–13,973. N.S. (Eds.), Tectonic Modelling: A Volume in Honor of Hans Ramberg: Geol. Soc.
Miyakawa, A., Yamada, Y., Matsuoka, T., 2010. Effect of increased shear stress along a Am. Mem., 193, pp. 179–190.
plate boundary fault on the formation of an out-of-sequence thrust and a break Schreurs, G., Buiter, S.J.H., Boutelier, D., Corti, G., Costa, E., Cruden, A.R., Daniel, J.-M.,
in surface slope within an accretionary wedge, based on numerical simulations. Hoth, S., Koyi, H.A., Kukowski, N., Lohrmann, J., Ravaglia, A., Schlische, R.W.,
Tectonophysics 484, 127–138. Withjack, M.O., Yamada, Y., Cavozzi, C., Delventisette, C., Elder Brady, J.A.,
Moore, J.C., Silver, E.A., 1987. Continental margin tectonics: submarine accretionary Hoffmann-Rothe, A., Mengus, J.-M., Montanari, D., Nilforoushan, F., 2006. Ana-
prisms. Reviews of Geophysics 25, 1305–1312. logue benchmarks of shortening and extension experiments. In: Buiter, S.J.H.,
Moresi, L., Mühlhaus, H.-B., 2006. Anisotropic viscous models of large-deformation Schreurs, G. (Eds.), Analogue and numerical modelling of crustal-scale pro-
Mohr–Coulomb failure. Philosophical Magazine 86, 3287–3305. doi:10.1080/ cesses,Geol. Soc. London Spec. Publ., 253, pp. 1–27.
14786430500255419. Schreurs, G., Buiter, S., Burberry, C., Callot, J.-P., Cavozzi, C., Cerca, M., Cristallini, E., Cruden,
Morgan, J.K., Boettcher, M.S., 1999. Numerical simulations of granular shear zones A., Chen, J.-H., Cruz, L., Daniel, J.-M., Garcia, V.H., Gomes, C., Grall, C., Guzmán, C.,
using the distinct element method 1. Shear zone kinematics and the micromecha- Hidayah, T.N., Hilley, G., Lu, C.-Y., Klinkmüller, M., Koyi, H., Macauley, J., Maillot, B.,
nics of localization. Journal of Geophysical Research 104, 2703–2719. Meriaux, C., Nilfouroushan, F., Pan, C.-C., Pillot, D., Portillo, R., Rosenau, M., Schellart,
Morgan, J.K., Karig, D.E., 1995. Kinematics and a balanced and restored cross-section W.P., Schlische, R., Take, A., Vendeville, B., Vettori, M., Vergnaud, M., Wang, S.-H.,
across the toe of the eastern Nankai accretionary prism. Journal of Structural Geol- Withjack, M., Yagupsky, D., Yamada, Y., 2010. Quantitative comparisons of analogue
ogy 17, 31–45. models of brittle wedge dynamics. Geophysical Research Abstracts 12, EGU2010-
Morley, C.K., King, R., Hillis, R., Tingay, M., Backe, G., 2011. Deepwater fold and thrust 11709-2.
belt classification, tectonics, structure and hydrocarbon prospectivity: a review. Schulze, D., 1994. Entwicklung und Anwendung eines neuartigen Ringschergerätes.
Earth-Science Reviews 104, 41–91. doi:10.1016/j.earscirev.2010.09.010. Aufbereitungstechnik 35, 524–535.
Mourgues, R., Cobbold, P.R., 2003. Some tectonic consequences of fluid overpressures Selzer, C., Buiter, S.J.H., Pfiffner, O.A., 2007. Sensitivity of shear zones in orogenic
and seepage forces as demonstrated by sandbox modelling. Tectonophysics 376, wedges to surface processes and strain softening. Tectonophysics 437, 51–70.
75–97. doi:10.1016/S0040-1951(03)00348-2. doi:10.1017/j.tecto.2007.02.020.
Mourgues, R., Cobbold, P.R., 2006. Thrust wedges and fluid overpressures: sandbox Selzer, C., Buiter, S.J.H., Pfiffner, O.A., 2008. Numerical modeling of frontal and basal ac-
models involving pore fluids. Journal of Geophysical Research 111, B05404. cretion at collisional margins. Tectonics 27, TC3001. doi:10.1029/2007TC002169.
doi:10.1029/2004JB003441. Simpson, G., 2006. Modelling interactions between fold-thrust belt deformation, fore-
Mühlhaus, H.B., Vardoulakis, I., 1987. The thickness of shear bands in granular mate- land flexure and surface mass transport. Basin Research 18, 125–143.
rials. Géotechnique 37, 271–283. doi:10.1111/j.1365-2117.2006.00287.x.
Mulugeta, G., 1988. Modelling the geometry of Coulomb thrust wedges. Journal of Simpson, G., 2009. Mechanical modelling of folding versus faulting in brittle-ductile
Structural Geology 10, 847–859. wedges. Journal of Structural Geology 31, 369–381.
Naylor, M., Sinclair, H.D., Willett, S., Cowie, P.A., 2005. A discrete element model for Simpson, G., 2010a. Formation of accretionary prisms influenced by sediment subduc-
orogenesis and accretionary wedge growth. Journal of Geophysical Research 110, tion and supplied by sediments from adjacent continents. Geology 38, 131–134.
B12403. doi:10.1029/2003JB002940. doi:10.1130/G30461.1.
Nieuwland, D.A., Leutscher, J.H., Gast, J., 2000. Wedge equilibrium in fold-and-thrust Simpson, G., 2010b. Influence of the mechanical behaviour of brittle–ductile fold-thrust
belts: prediction of out-of-sequence thrusting based on sandbox experiments belts on the development of foreland basins. Basin Research 22, 139–156.
and natural examples. Geologie en Mijnbouw 79, 81–91. doi:10.1111/j.1365-2117.2009.00406.x.
Nilforoushan, F., Koyi, H.A., 2007. Displacement fields and finite strains in a sandbox Simpson, G., 2011. Mechanics of non-critical fold-thrust belts based on finite element
model simulating a fold-thrust-belt. Geophysical Journal International 168, models. Tectonophysics 499, 142–155. doi:10.1016/j.tecto.2011.01.004.
1341–1355. doi:10.1111/j.1365-246X.2007.03341.x. Smart, K.J., Couzens-Schultz, B.A., 2001. Mechanics of blind thrusting: comparison of
Nilforoushan, F., Koyi, H.A., Swantesson, J.O.H., Talbot, C.J., 2008. Effect of basal friction numerical and physical modelling. Journal of Geology 109, 771–779.
on surface and volumetric strain in models of convergent settings measured by Smit, J.H.W., Brun, J.P., Sokoutis, D., 2003. Deformation of brittle–ductile thrust wedges
laser scanner. Journal of Structural Geology 30, 366–379. doi:10.1016/ in experiments and nature. Journal of Geophysical Research 108. doi:10.1029/
j.jsg.2007.09.013. 2002JB002190.
Panian, J., Wiltschko, D., 2007. Ramp initiation and spacing in a homogeneous thrust Stockmal, G.S., Beaumont, C., Nguyen, M., Lee, B., 2007. Mechanics of thin-skinned fold-
wedge. Journal of Geophysical Research 112, B05417. doi:10.1029/2004JB003596. and-thrust belts: insights from numerical models. In: Sears, J.W., Harms, T.A.,
Panien, M., Schreurs, G., Pfiffner, A., 2006. Mechanical behaviour of granular materials Evenchick, C.A. (Eds.), Whence the Mountains? Inquiries into the Evolution of
used in analogue modelling: insights from grain characteristics, ring-shear tests Orogenic Systems: A Volume in Honor of Raymond A. Price, Geol. Soc. Am. Spec.
and analogue experiments. Journal of Structural Geology 28, 1710–1724. Paper, 433, pp. 63–98. doi:10.1130/2007.2433(04).
doi:10.1016/j.jsg.2006.05.004. Storti, F., McClay, K., 1995. Influence of syntectonic sedimentation on thrust wedges in
Persson, K.S., 2001. Effective indenters and the development of double-vergent oro- analogue models. Geology 23, 999–1002.
gens: insights from analogue sand models. In: Koyi, H.A., Mancktelow, N.S. Storti, F., Salvini, F., McClay, K., 2000. Synchronous and velocity-partitioned thrusting
(Eds.), Tectonic Modelling: A Volume in Honor of Hans Ramberg: Geol. Soc. A. and thrust polarity reversal in experimentally produced, doubly-vergent thrust
Mem., 193, pp. 191–206. wedges: implications for natural orogens. Tectonics 19, 378–396.
Persson, K.S., Sokoutis, D., 2002. Analogue models of orogenic wedges controlled by Strayer, L.M., Hudleston, P.J., Lorig, L.J., 2001. A numerical model of deformation and
erosion. Tectonophysics 356, 323–336. fluid-flow in an evolving thrust wedge. Tectonophysics 335, 121–145.
Pfiffner, O.A., 2006. Thick-skinned and thin-skinned styles of continental contraction. Teixell, A., Koyi, H.A., 2003. Experimental and field study of the effects of lithological
In: Mazzoli, S., Butler, R.W.H. (Eds.), Styles of continental contraction: Geol. Soc. contrasts on thrust-related deformation. Tectonics 22. doi:10.1029/2002TC001407.
Am. Spec. Paper, 414, pp. 153–177. doi:10.1130/2006.2414(09). Vardoulakis, I., 1980. Shear band inclination and shear modulus of sand in biaxial tests.
Poblet, J., Lisle, R.J., 2011. Kinematic evolution and structural styles of fold-and-thrust belts. International Journal for Numerical and Analytical Methods in Geomechanics 4,
Geological Society of London—Special Publication 349, 1–24. doi:10.1144/SP349.1. 103–119.
Popov, A.A., Sobolev, S.V., 2008. SLIM3D: a tool for three-dimensional thermomechani- Vermeer, P.A., 1990. The orientation of shear bands in biaxial tests. Géotechnique 40,
cal modeling of lithospheric deformation with elasto-visco-plastic rheology. Phys- 223–236.
ics of the Earth and Planetary Interiors 171, 55–75. doi:10.1016/j.pepi.2008.03.007. Von Huene, R., Scholl, D.W., 1991. Observations at convergent margins concerning sed-
Price, R.A., 1981. The Cordilleran foreland thrust and fold belt in the southern Canadian iment subduction, subduction erosoin, and the growth of continental crust. Re-
Rocky Mountains. In: McClay, K.R., Price, N.J. (Eds.), Thrust and nappe tectonics: views of Geophysics 29, 279–316.
Geol. Soc. Lond., Spec. Publ., 9, pp. 427–448. Wang, K., Hu, Y., 2006. Accretionary prisms in subduction earthquake cycles: the theo-
Richard, P., Krantz, R.W., 1991. Experiments on fault reactivation in strike-slip mode. ry of the dynamic Coulomb wedge. Journal of Geophysical Research 111, B06410.
Tectonophysics 188, 117–131. doi:10.1029/2005JB004094.
S.J.H. Buiter / Tectonophysics 530-531 (2012) 1–17 17

Wang, W.M., Sluys, L.J., de Borst, R., 1996. Interaction between material length scale Willett, S.D., Brandon, M.T., 2002. On steady states in mountain belts. Geology 30,
and imperfection size for localisation phenomena in viscoplastic media. European 175–178.
Journal of Mechanics—A/Solids 15, 447–464. Willett, S., Beaumont, C., Fullsack, P., 1993. Mechanical model for the tectonics of dou-
Wang, K., He, J., Hu, Y., 2006. A note on pore fluid pressure ratios in the Coulomb wedge bly vergent compressional orogens. Geology 21, 371–374.
theory. Geophysical Journal International 33, L19310. doi:10.1029/2006GL027233. Willis, B., 1893. The mechanics of Appalachian structures. US Geological Survey 13th
Westbrook, G.K., Smith, M.J., Peacock, J.H., Poulter, M.J., 1982. Extensive underthrusting Annual Report, Part 2, pp. 211–281.
of undeformed sediment beneath the accretionary complex of the Lesser Antilles Withjack, M.O., Jamison, W.R., 1986. Deformation produced by oblique rifting. Tectono-
subduction zone. Nature 300, 625–628. physics 126, 99–124.
Westbrook, G.K., Ladd, J.W., Buhl, P., Bangs, N., Tilley, G.J., 1988. Cross section of an ac- Yamato, P., Kaus, B.J.P., Mouthereau, F., Castelltort, S., 2011. Dynamic constraints on the
cretionary wedge: Barbados Ridge complex. Geology 16, 631–635. crustal-scale rheology of the Zagros fold belt, Iran. Geology 39, 815–818.
Willett, S.D., 1992. Dynamic and kinematic growth and change of a Coulomb wedge. In: doi:10.1130/G32136.1.
McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall, New York, pp. 19–31. Zhao, W.-L., Davis, D.M., Dahlen, F.A., Suppe, J., 1986. Origin of convex accretionary
Willett, S.D., 1999a. Orogeny and orography: the effects of erosion on the structure of wedges: evidence from Barbados. Journal of Geophysical Research 91, 10,246–10,258.
mountain belts. Journal of Geophysical Research 104, 28,957–28,981.
Willett, S.D., 1999b. Rheological dependence of extension in wedge models of conver-
gent orogens. Tectonophysics 305, 419–435.

You might also like