journal of structural geology · journal of structural geology 87 (2016) 134e143 submarine...

10
Structural consequences of cohesion in gravitational instabilities triggered by uid overpressure: Analytical derivation and experimental testing R. Mourgues a, * , A.C.G. Costa b, c , F.O. Marques d , A. Lacoste e , A. Hildenbrand c a L.P.G., UMR-CNRS 6112, Universit e du Maine, 72089 Le Mans, France b Universidade de Lisboa and IDL, Lisboa, Portugal c GEOPS, Univ. Paris-Sud, CNRS, Universit e Paris-Saclay, Rue du Belv ed ere, B^ at. 504-509, 91405 Orsay, France d Universidade de Lisboa, Lisboa, Portugal e GeHCO, E.A.6293, Universit e de Tours, Tours, France article info Article history: Received 12 October 2015 Received in revised form 28 March 2016 Accepted 6 April 2016 Available online 13 April 2016 Keywords: Gravitational spreading Landslide Effect of cohesion Fluid overpressure Critical taper theory Listric faults abstract The critical taper theory of Coulomb wedges has been classically applied to compressive regimes (accretionary prisms/fold-and-thrust belts), and more recently to gravitational instabilities. Following the initial hypothesis of the theory, we provide an alternative expression of the exact solution for a non- cohesive wedge by considering the balance of forces applied to the external surfaces. Then, we use this approach to derive a solution for the case of cohesive wedges. We show that cohesion has conspicuous structural effects, including a minimum length required for sliding and the formation of listric faults. The stabilizing effect of cohesion is accentuated in the foremost thin domain of the wedge, dening a required Minimum Failure Length (MFL), and producing sliding of a rigid mass above the detachment. This MFL decreases with less cohesion, a smaller coefcient of internal friction, larger uid overpressure ratio, and steeper upper and basal surfaces for the wedge. Listricity of the normal faults depends on the uid overpressure magnitude within the wedge. For moderate uid overpressure, normal faults are curved close to the surface, and become straight at depth. In contrast, where uid overpressure exceeds a critical value corresponding to the uid pressure required to destabilize the surface of a noncohesive wedge, the state of stress changes and rotates at depth. The faults are straight close to the surface and listric at depth, becoming parallel to the upper surface if the wedge is thick enough. We tested some of these structural effects of a cohesive wedge on gravitational instabilities using analogue models where cohesive material was subjected to pore-uid pressure. The shape of the faults obtained in the models is consistent with the predictions of the theory. © 2016 Elsevier Ltd. All rights reserved. 1. Introduction Events of large-scale gravitational destabilization occur in a wide range of geological settings (e.g., Moore et al., 1989; Haidason et al., 2004; Lacoste et al., 2009). Such events result from progressive changes in physical parameters that are critical for the stability state of a given slope (e.g., decrease of the material's strength and surface morphology due to weathering/erosion, sedimentation, tectonics, changes to pore-pressure conditions) (e.g., Voight and Elsworth, 1997; Masson et al., 2006). The identication of the factors lead- ing to stability decrease matters greatly to geosciences hazard analysis with respect to society and the economy (e.g., geotechnics and civil engineering, volcanology, oil industry). However, in many cases, only the surface deformation or the scar/deposits are detectable (e.g., Hildenbrand et al., 2012; Costa et al., 2014). Geophysical methods characterize the geometry of the basal detachment and nature of the internal deformation of the displaced volume (Morgan et al., 2003; Haidason et al., 2004). Geomechanical tests also provide helpful information, but sam- pling is sometimes difcult or impossible. Detailed structural analysis can be used to estimate indirectly some of the mechanical characteristics of the rocks involved in the unstable mass. Alter- natively, efcient mathematical models are required to describe the various structural outcomes from the destabilization. Among the destabilizing factors, high pore-uid pressure is often invoked for causing low-dip slopes in sub-aerial and * Corresponding author. E-mail address: [email protected] (R. Mourgues). Contents lists available at ScienceDirect Journal of Structural Geology journal homepage: www.elsevier.com/locate/jsg http://dx.doi.org/10.1016/j.jsg.2016.04.008 0191-8141/© 2016 Elsevier Ltd. All rights reserved. Journal of Structural Geology 87 (2016) 134e143

Upload: others

Post on 17-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

lable at ScienceDirect

Journal of Structural Geology 87 (2016) 134e143

Contents lists avai

Journal of Structural Geology

journal homepage: www.elsevier .com/locate/ jsg

Structural consequences of cohesion in gravitational instabilitiestriggered by fluid overpressure: Analytical derivation andexperimental testing

R. Mourgues a, *, A.C.G. Costa b, c, F.O. Marques d, A. Lacoste e, A. Hildenbrand c

a L.P.G., UMR-CNRS 6112, Universit�e du Maine, 72089 Le Mans, Franceb Universidade de Lisboa and IDL, Lisboa, Portugalc GEOPS, Univ. Paris-Sud, CNRS, Universit�e Paris-Saclay, Rue du Belv�ed�ere, Bat. 504-509, 91405 Orsay, Franced Universidade de Lisboa, Lisboa, Portugale GeHCO, E.A.6293, Universit�e de Tours, Tours, France

a r t i c l e i n f o

Article history:Received 12 October 2015Received in revised form28 March 2016Accepted 6 April 2016Available online 13 April 2016

Keywords:Gravitational spreadingLandslideEffect of cohesionFluid overpressureCritical taper theoryListric faults

* Corresponding author.E-mail address: [email protected] (R

http://dx.doi.org/10.1016/j.jsg.2016.04.0080191-8141/© 2016 Elsevier Ltd. All rights reserved.

a b s t r a c t

The critical taper theory of Coulomb wedges has been classically applied to compressive regimes(accretionary prisms/fold-and-thrust belts), and more recently to gravitational instabilities. Followingthe initial hypothesis of the theory, we provide an alternative expression of the exact solution for a non-cohesive wedge by considering the balance of forces applied to the external surfaces. Then, we use thisapproach to derive a solution for the case of cohesive wedges. We show that cohesion has conspicuousstructural effects, including a minimum length required for sliding and the formation of listric faults. Thestabilizing effect of cohesion is accentuated in the foremost thin domain of the wedge, defining arequired Minimum Failure Length (MFL), and producing sliding of a rigid mass above the detachment.This MFL decreases with less cohesion, a smaller coefficient of internal friction, larger fluid overpressureratio, and steeper upper and basal surfaces for the wedge. Listricity of the normal faults depends on thefluid overpressure magnitude within the wedge. For moderate fluid overpressure, normal faults arecurved close to the surface, and become straight at depth. In contrast, where fluid overpressure exceeds acritical value corresponding to the fluid pressure required to destabilize the surface of a noncohesivewedge, the state of stress changes and rotates at depth. The faults are straight close to the surface andlistric at depth, becoming parallel to the upper surface if the wedge is thick enough. We tested some ofthese structural effects of a cohesive wedge on gravitational instabilities using analogue models wherecohesive material was subjected to pore-fluid pressure. The shape of the faults obtained in the models isconsistent with the predictions of the theory.

© 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Events of large-scale gravitational destabilization occur in awiderange of geological settings (e.g.,Moore et al.,1989;Haflidason et al.,2004; Lacoste et al., 2009). Such events result from progressivechanges in physical parameters that are critical for the stability stateof a given slope (e.g., decrease of thematerial's strength and surfacemorphology due to weathering/erosion, sedimentation, tectonics,changes to pore-pressure conditions) (e.g., Voight and Elsworth,1997; Masson et al., 2006). The identification of the factors lead-ing to stability decrease matters greatly to geosciences hazard

. Mourgues).

analysis with respect to society and the economy (e.g., geotechnicsand civil engineering, volcanology, oil industry).

However, in many cases, only the surface deformation or thescar/deposits are detectable (e.g., Hildenbrand et al., 2012; Costaet al., 2014). Geophysical methods characterize the geometry ofthe basal detachment and nature of the internal deformation of thedisplaced volume (Morgan et al., 2003; Haflidason et al., 2004).Geomechanical tests also provide helpful information, but sam-pling is sometimes difficult or impossible. Detailed structuralanalysis can be used to estimate indirectly some of the mechanicalcharacteristics of the rocks involved in the unstable mass. Alter-natively, efficient mathematical models are required to describe thevarious structural outcomes from the destabilization.

Among the destabilizing factors, high pore-fluid pressure isoften invoked for causing low-dip slopes in sub-aerial and

Page 2: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143 135

submarine structures at various scales: e.g., large-scale flank col-lapses in volcanic islands (Iverson, 1995; Day, 1996), sub-aeriallandslides in active margins associated with river incision (Waita-whiti landslides, New Zealand e Lacoste et al., 2009, 2011), andsubmarine landslides affecting passive margins (Amazon Fan e

Cobbold et al., 2004; Mourgues et al., 2009; Gulf of Mexico -Flemings et al., 2008; Storegga slide/North Sea Fane Kvalstad et al.,2005b). Fluid overpressure is generated through several mecha-nisms, such as compaction disequilibrium (Walder and Nur, 1984;Day, 1996), mechanical and thermal pressurization of the porousmedium (e.g., thermal and mechanic effects of magmatic intrusion,and associated effect of magmatic degassing e Iverson, 1995; Day,1996; Elsworth and Voight, 1996; Voight and Elsworth, 1997;Reid, 2004), seismic loading (Elsworth and Voight, 1996), chang-ing sea level (Iverson, 1995; Quidelleur et al., 2008; Smith et al.,2013), groundwater flow of meteoric water (Reid, 2004), andchemical reactions as in hydrocarbon generation (e.g., Cobboldet al., 2004, 2013; Zanella et al., 2014a, 2014b).

The critical taper model developed by Dahlen and co-authors isone of the most commonly used mathematical models for inter-preting deformation systems having pore-fluid pressure (Daviset al., 1983; Dahlen et al., 1984; Dahlen, 1984, 1990). It has beenwidely applied to studying the shape of accretionary wedges andthrust belts, and estimating the weakness of basal detachments,often as a function of fluid pressure. Awedge of a Coulombmaterialsubjected to lateral compression will deform internally until acritical taper forms. With this state, the prism and the basaldetachment are everywhere on the verge of shear failure. Then,deformation will proceed with the growth of the wedge, includingthe self-similar growth for non-cohesive materials, through ac-cretion of new material at the toe that maintains the critical taperand sliding along the basal detachment (Davis et al., 1983; Dahlen,1984; Dahlen et al., 1984). The mathematical formulation of thecritical taper has multiple solutions, depending on the shearorientation on the basal detachment and the stress state within thewedge (i.e., contractional or extensional).

Mourgues et al. (2014) pointed out that few studies focused onthe applicability of the solution to gravitational spreading andgliding along passive margins, where elevated pore-fluid pressureis common in sediments, and where numerous gravitationalstructures, such as landslides and debris flows occur. Mourgueset al. (2014) investigated the applicability of the solution to gravi-tational instabilities for non-cohesive material with high pore-fluidpressure, in the absence of any external compressive or extensionalforces. Mourgues et al. (2014) also proposed a mathematicalexpression differing fromDahlen (1984) for structural systems withgravitational instabilities, and they verified the predictions of theanalytical model for slope instabilities by using physical experi-ments where compressed air was applied at the base of dry sandwedges to trigger gravitational instabilities (Mourgues andCobbold, 2006a, 2006b; Lacoste et al., 2012). In their analysis,Mourgues et al. (2014) neglected the cohesion of the wedge, whichwas consistent with the prior work of Davis et al. (1983) and Dahlen(1984). However, cohesion is a critical parameter for rock strength,allowing for steep slopes and vertical cliffs to be stable, andtherefore ideally should not be neglected in stability analysis (e.g.,del Potro et al., 2013). In submarine settings, some landslidesgenerate large blocks of undisrupted material, reflecting the initialcohesion of the sediments. Two famous major landslides presentsuch characteristics: (1) the Storegga slide affected a sedimentvolume of 2400e3200 km3 in the continental platform of the NorthSea, with lateral spreading of large scale blocks (Toreva blocks)along 1.1e1.4� dipping failure surfaces (Haflidason et al., 2004; Brynet al., 2005; Kvalstad et al., 2005a, 2005b); and (2) the Nu'uanudebris avalanche affected a volume of 2500e3500 km3 in the NE of

O'ahu Island in Hawaii, and involved the transport of a 30-km longand 600-km3 block (the Tuscaloosa block) along ca. 55 km (Mooreet al., 1989; Moore and Clague, 2002; Satake et al., 2002). The as-sociation of these major collapse events to fluid overpressure(Iverson, 1995; Kvalstad et al., 2005a) highlights the interest ofstudying the occurrence of gravitational destabilization assisted byfluid overpressure in cohesive materials.

Cohesion was first introduced in the critical taper model byDahlen et al. (1984). These authors predicted that cohesion has astrong effect in the thinnest part of the wedge, while the defor-mation in the thickest part is similar to that of non-cohesive ma-terial (Davis et al., 1983; Dahlen et al., 1984). This implies that theupper surface of a Coulomb wedge will be concave: flatter in theforemost thinner sector more resistant to deformation, and steeperin the thicker sector where the slope tends towards that expectedfor a non-cohesive wedge (Davis et al., 1983).

Following the initial hypothesis of the Coulomb critical tapertheory, we reformulate the analytical approach presented inMourgues et al. (2014) to study the occurrence of gravitationaldestabilization in a cohesive wedge. We show that cohesion hasconspicuous structural effects, including a minimum length ofsliding and the formation of listric faults. This last prediction is thenverified in the present study by physical experiments involvingpore-fluids in cohesive and permeable granular materials.

2. The non-cohesive critical taper e a force equilibriumapproach

The critical taper theory is based on the assumption that theinternal state of stress of an homogeneous wedge composed ofmaterial deforming according to the Mohr-Coulomb criterion is onthe verge of failure everywhere (Davis et al., 1983). The shape of thewedge, growing self-similarly, depends on the strength of thematerial, and on the relative magnitude of the basal friction.

2.1. Solutions derived from the equilibrium of internal stresses

Dahlen (1984) provided an exact solution for non-cohesivewedges by expressing the total taper angle of the critical wedge(expressed as the sum of the critical surface slope, a, and that of thebasal detachment, b, Eq. (1)) as a function of pore pressure ratios(considered as constant) along the basal detachment (lb) andwithin the wedge (l), coefficients of sliding friction along the basaldetachment (mb) and within the wedge (m):

aþ b ¼ Jo þJb (1)

where Jo and Jb are the angles between the maximum principalstress and the upper surface or the basal detachment, respectively(Fig. 1a). These two angles are written more explicitly:

Jo ¼ 0:5 arcsin�sin a0

sin f

�� 0:5a0 (2)

with

a0 ¼ arctan

�1

1� l*tan a

�(3)

Jb ¼ 0:5 arcsin�sin f0

bsin f

�� 0:5f0

b (4)

With f the angle of friction of the wedge and f0b the effective

angleof frictionon thedetachment, anddefinedbyDahlen (1984)as:

Page 3: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

Fig. 1. a. Cross section of a wedge, showing the cartesian coordinates systems and variables used in the analysis (including Jo and Jb angles as used in Dahlen, 1984). b. Forcesconsidered in the assessment of the wedge's stability along the basal detachment.

Fig. 2. Critical taper solutions for non-cohesive wedge given by equation (6). Thedashed line represents the solution for a compressive state of stress. The black boldline represents the solution for a gravity-driven state of stress that yields extensionalbehavior. The grey area shows domains of slope instability triggered by gravity only.Gravitational spreading requires the detachment surface to be subjected to a minimalpore-fluid pressure (l*b

m). The line FS ¼ 1 limits the area of shallow slides.

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143136

m0b ¼ tan f0b ¼ 1� l*b

1� l*mb (5)

With l* the pore pressure ratio:

l* ¼ Povr0gz cos a

(6)

where r0 is the density corrected for hydrostatic buoyancy(r0 ¼ r� rwwith r the bulk density and rw the density of water), g isthe gravitational acceleration, and Pov is the fluid overpressure(Hubbert and Rubey, 1959). Pov is defined as the excess of porepressure with reference to the hydrostatic pressure at the samedepth (Mourgues and Cobbold, 2006b).

l* is defined within the wedge and lb* is the pore-fluid over-

pressure ratio on the basal detachment.Mourgues et al. (2014) provided an alternative formulation of

Dahlen's expression (Dahlen, 1984) and Lehner's graphical solution(Lehner, 1986). Mourgues et al. (2014) expressed the fluid over-pressure ratio (or the basal friction) required for the wedge to slideon a basal low-resistance layer as:

l*b ¼ 1��1� l*

� E2mbE1

(7)

with

E1 ¼�1� l*

�fY þ ð1� YÞcos½2ða� bÞ�g þ tanðaÞsin½2ða� bÞ�

(8)

E2 ¼�1� l*

�fðY � 1Þsin½2ða� bÞ�g þ tanðaÞcos½2ða� bÞ�

(9)

Y ¼ 1� sin fffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1� FS2

p

cos2 ffor a tensile state of stress (10)

Y¼1þsinfffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1�FS2

p

cos2ffor a compressional state of stress (11)

and

FS ¼ tan a�1� l*

�tan f

(12)

This solution is better suited to the study of slope instabilities(Mourgues et al., 2014), and it provides important indications on

the stability of the surface slope and the potential triggering ofshallow slumps through the expression of the factor of safety FS(Eq. (12), and Fig. 2).

2.2. Solution derived from the balance of external forces applied ona triangular segment of the wedge

Equations (1) and (7) were derived from the equilibrium of in-ternal stresses within the wedge. An alternative solution is toconsider the balance of external forces applied on a triangularsegment of the wedge. Dahlen used this approach in his reviewpaper (Dahlen, 1990) to introduce the critical taper theory througha more intuitive approach, but he used a simplified state of stressesand did not derive the exact solution this way. In such an approach,the balance of forces applied on a triangular segment of length l andacting in the x1 direction (Fig. 1b) is:

Fs þ Fx1w þ Fx1S þ Fb ¼ 0 (13)

Fs is the force exerted by sx1x1 on the sidewall of thewedge. Fwx1 isthe x1 component of the gravitational body force. Fsx1 is the net x1component of the force resulting from the pore-fluid pressure(seepage force), and Fb is the frictional force exerted at the base.Detailed expressions of these forces are given in Appendix A.

An exact solution can be found for a non-cohesive wedge, by

Page 4: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143 137

considering the static equilibrium of the system defined by Dahlen(1984) and expressions derived by Mourgues et al. (2014):

l*b ¼ 1� 1� l*

E1mb

��E3 þ l*

�tanða� bÞ þ sin b

cos a cosða� bÞ�

(14)

with

E3 ¼�1� l*

�ðY þðY �1Þcos 2ða�bÞÞ� tan a sin 2ða� bÞ (15)

Solutions 1, 7 and 13 are numerically identical. Wang et al.(2006), Mourgues et al. (2014) and Yuan et al. (2015) pointed outthat the use of an effective angle of friction f’b (Eq. (5)) introducedby Dahlen is wrong, due to an error in the reference coordinateused to define l. Mourgues et al. (2014) proposed a corrected valueof the effective basal friction:

m0b ¼ mb

1þ l* � l*b

E1

!(16)

and obtained an alternative solution for the critical taper. Using thiscorrected friction, Eq. (14) becomes:

l*b ¼ E1 þ l* � 1mb

��E3 þ l*

�tanða� bÞ þ sin b

cos a cosða� bÞ�

(17)

3. Gravitational instabilities in a cohesive wedge, andstructural consequences

In the model derived from external forces (Section 2.2), cohesionwas neglected. By introducing a significant cohesion within thewedge, Fs is modified (see Appendix A). It also slightly changes Fb bymodifying szz. Cohesion induces non-linearity and a simple exact so-lution can no longer be found. Consequently, Fs has to be numericallyintegrated and the solution of the force balance is iteratively found.

3.1. Minimal Failure Length (MFL)

On one hand, in a non-cohesive wedge, sliding along a low-resistance detachment plane can be triggered as soon as the grav-itational potential of the system is large enough to overcome thebasal frictional resistance. This behavior only requires the detach-ment surface to be subjected to a minimal pore-fluid pressure l*b

m

(Fig. 2). This pressure value (Eq. (17)) does not depend on the lengthof the wedge, therefore driving and resisting forces are identicaleverywhere within the wedge (Figs. 3a and 4a), and normal faultsmay appear everywhere (Fig. 3a).

On the other hand, cohesion allows steeper stable slopes, byincreasing the shear strength of the material. Fs acts as a resistingforce near the surface, and becomes a driving force at depth (Fig. 4b).Despite the wedge being considered everywhere on the verge offailure, the equilibrium of forces requires a Minimal Failure Length(MFL), as illustrated in Fig. 4b and c. The plot of the driving andresisting forces along the basal detachment (Fig. 3b) shows that forwedge lengths (L) smaller than theMFL, thedriving force Fd is smallerthan the resisting force Fb and the wedge remains stable along itsbasal detachment. In other words, no gravitational instabilitiesrooting on the basal detachment can develop in wedges havinglengths shorter than the MFL. For wedges longer than the MFL, thefirst normal faults to form root on the basal detachment at aminimaldistance from the toe (at least at theMFL distance). TheMFL varies asa function of several factors. It increases with increasing cohesionand frictionvalues in thewedgeandalong thebasaldetachment, and

it decreases with increasing pore pressure (l* and l*b). Therefore, ifthe pore pressure progressively increases, then the MFL becomessmaller than the length of the wedge L. Formation of the normalfaults will occur, at least at an MFL distance from the front of thewedge,while the frontal/thinner segmentof thewedge (L<MFL)willslide along the basal detachment without internal deformation(Fig. 3b). This is an important difference with non-cohesive wedges,where deformationmay start everywhere in thewedge (Figs. 3a and4a). The necessaryMFL and the formation of a sliding mass with nomajor internal deformation are some aspects that are also observedin slides having a downslope buttress (Fig. 4c, Mourgues andCobbold, 2006a; Mourgues et al., 2009). The presence of such adownslope buttress adds a resisting force similar to the effect ofcohesion in the wedge. The sliding mass then comprises threedistinct domains: an extensional domain upslope, a non-deformedslab at mid-slope, and a contractional domain downslope(Mourgues andCobbold, 2006a;Mourgues et al., 2009; Lacoste et al.,2012).

3.2. Conditions for shallow sliding

In non-cohesive wedges, if pore pressure within the wedgebecomes too large, the factor of safety FS reaches values of 1 (Eq.(12)) and the surface becomes unstable, thus triggering shallowslides. Detachments form parallel to the surface at any depth(Fig. 5a). The critical pore pressure ratio (l*ncc Þ required to triggersuch shallow sliding in non-cohesive material can easily be found(Hubbert and Rubey, 1959; Mandl and Crans, 1981) by consideringfailure on a plane parallel to the surface:

s0xz ¼ ms0zz (18)

Replacing s0zz and s0xz with expressions A1 and A2 (see AppendixA) leads to:

l*ncc ¼ 1� tan a

tan f(19)

Given that cohesion (Co) increases the stability of the surfaceslope, we add cohesion to the failure criterion (Eq. (18)) and thecritical pore pressure ratio becomes:

l*c ¼ 1� tan a

tan fþ Cor0gz cosa tan f

¼ l*ncc þ Cor0gz cosa tan f

(20)

Equation (20) shows that l*c increases with Co. Where l*<l*cnc,

the cover remains stable regardless of the depth. Where l*>l*cnc, a

detachment may form at depth Zc (Fig. 5b):

zc ¼ Co�l* � l*ncc

�rg cos a tan f

(21)

This critical depth varies with cohesion and pore pressure l*

(Fig. 5c). When the pore pressure increases, slides may form closerto the surface. If the wedge is thinner than Zc, it remains internallystable (Fig. 6b). At the rear of the wedge, where the thickness be-comes larger than Zc, a new detachment may form parallel to thesurface and may promote the rooting of normal faults (Fig. 6c).

3.3. Shape of the normal faults

In the non-cohesive critical taper theory, the angle jo betweenthe maximum principal stress and the surface is constant throughthe wedge, and therefore normal faults are predicted to be straight.Their dip depends on the magnitude of pore-fluid overpressure l*,as argued by Mourgues and Cobbold (2003). More generally,

Page 5: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

Fig. 3. Balance between driving forces Fd ¼ Fs þ Fw þ Fs and resisting forces Fb for a non-cohesive wedge satisfying equation (16) (a) and for a cohesive wedge requiring a MinimalFailure Length (MFL) (b).

Fig. 4. Representation of the forces acting within a wedge (black), and respectivepatterns of deformation (grey) in the case of: (a) non-cohesive wedge, (b) cohesivewedge, (c) buttressed cohesive wedge.

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143138

Mourgues and Cobbold (2003) showed that the dips of normalfaults in a sloping sedimentary sequence are controlled by theseepage forces induced by the pore-fluid overpressure gradient.They also showed that listric normal faults can form wherepermeability and pressure gradient (or l*) vary with depth.

In the case of a cohesive wedge, the non-linearity resulting fromcohesion induces stress rotation and therefore listric faults, withoutany variation of l* with depth.

Fig. 7 shows the orientation of the faults relatively to the z-axis,as a function of depth z, determined using the following expres-sions from Mourgues and Cobbold (2003):

tanð2uÞ ¼ �2s0xz

s0xx � s0zz�

(22)

and

g ¼ 45� f=2 (23)

where u is the angle between s1 and the z-axis, and g is the anglebetween s1 and the generated fault. In these expressions, the valuesof s0zz, s0xz and s0xx are calculated with equation (A1), A2 and A12,respectively (see Appendix A).

Two cases must be distinguished, depending on the value of l*

and on the evolution of stresses with depth z (Figs. 6a and 7):

(1) Where l*<l*cnc, stress rotation occurs mainly close to the

surface, where the faults are steeper. At depth, where theinfluence of cohesion decreases, the orientation of themaximum principal stress tends towards jo and the faultsbecome straight (Figs. 6b and 7).

(2) Where l*>l*cnc, the faults have a listric shape. This shape

depends on the critical depth Zc, where faults become par-allel to the upper surface of the wedge (Figs. 6c and 7).

To test these last predictions, we used physical experimentsinvolving pore-fluids in cohesive and permeable granularmaterials.

4. Analogue modelling

4.1. Experimental setup and procedure

Fluid overpressure has been successfully modelled through in-jection of compressed air in experimental studies (e.g., Cobbold andCastro, 1999; Cobbold et al., 2001, Mourgues and Cobbold, 2003,2006a, 2006b; Mourgues et al., 2009; Lacoste et al., 2011, 2012)).Mourgues et al. (2014) were able to verify the predictions of thecritical Coulomb wedge theory in the case of gravitational in-stabilities by performing scaled experiments with pore pressureunder conditions close to the critical taper hypothesis. The maindifficulty was that l* and lb

* must be constant within the wedge andalong the detachment surface, respectively. In their experiments,the air pressure loss within the triangular edge of the model(Fig. 5b) had to be balanced by adding glass microbeads underneath(Mourgues et al., 2014).

However, even with such an apparatus, the basal pressure wasnot strictly constant and uncertainties on lb

* were strong (±0.1 forlb* ). For practical reasons, and because these uncertainties may arisewith the use of cohesive material, we thought that it was notrelevant to try to verify all the predictions of the cohesive model

Page 6: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

Fig. 5. a. Inanon-cohesive system, the failureparallel to theupper surface canoccurat anydepth;b. Ina cohesivesystem, the failureparallel to theupper surfacewilloccur fordepthsequalor superior to a critical value (Zc); c. The critical depth Zc decreases exponentially with increasing l*, and ismore easily attained in the case of steeper wedges (>a) and smaller cohesions.

Fig. 6. a. Representation of the Mohr diagram for l* below, equal or above a critical l* value (lc*). b. Where l* < lc*, Zc is not attained within the wedge, and the faults do not become

parallel to the upper surface of the wedge. c. Where l* > lc*, Zc is attained within the wedge, and the faults become parallel to the upper surface of the wedge.

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143 139

with such an experimental device.We therefore chose to verify onlypart of the predictions. Within the wedge itself, pore-fluid isobarsare supposed to be parallel to the surface. This pore pressure con-dition can be easily applied on an analoguemodel. Thus, we focusedon the stability of the surface and shape of normal faults.

In Mourgues and Cobbold (2003), extensional tests were per-formed in tilted tabular models of homogeneous non-cohesivematerials (fine sand), with upward injection of compressed air toverify the structural consequences of pore-fluid pressures andseepage forces. Here, we report on similar tests performed using

Fig. 7. Variation with depth of the angle between the z-axis and the faults for variousl* values, with l*c

nc ¼ 0.53. Notice the different shapes of the curves for l* higher andsmaller than l*c

nc. For l* < l*cn, faults are curve close to the surface and become straight

at depth. For l* > l*cn, faults flatten with depth becoming parallel to the surface at the

corresponding value of Zc.

cohesive materials, to verify the formation of listric faults predictedby the analytical model (Section 3.3). The experimental setup wassimilar to that used in Mourgues and Cobbold (2003). We built amodelmade of cohesivematerial resting on two overlapping sieves.Beneath the sieves was a pressure chamber that provided a uniformfluid pressure at the base of the model. By slowlymoving one of thesieves, we created a velocity discontinuity that induced the for-mation of normal faults (see Mourgues and Cobbold, 2003; for acomplete description of the apparatus). The cohesive material wascomposed of fine glass microbeads. The cohesion (Co¼ 140 ± 20 Pa)and angle of internal friction (f ¼ 26�) were determined with aseries of direct shear tests on the material compacted tor ¼ 1600 kg/m3. The model was 5-cm thick. The microbeads layerswere intercalated with dark thin layers of Silicon carbide, whichserved asmarkers of deformation. Themodel was then tilted 21�, sothat Zc was close to the base of the model for reasonable values ofl*, avoiding the risk of explosion due to high pore pressure(Mourgues and Cobbold, 2006a).

Three different pore pressures were imposed at the base of themodels: 0 Pa, 200 Pa and 400 Pa. These pressures correspond to l*

values of 0, 0.25 ± 0.1 and 0.5 ± 0.1, respectively. Considering theuncertainties for the thickness and density of themodel, and for themeasurement of pore pressure, we estimated l*with an uncertaintyof 0.1. Assuming Co ¼ 140 Pa, we estimated l*c

nc close to 0.21 andl*c ¼ 0.6 to activate a detachment along the base of the model.

4.2. Results and comparison with analytical predictions

In Fig. 8, we present the interpreted photographs taken duringthe experiments for the three different pore pressures. For l* ¼ 0

Page 7: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

Fig. 8. Interpreted lateral photographs illustrating cross-sectional views of analogue models, with assumed material cohesion of 140 Pa, and l* values of 0, 0.25 ± 0.1 and 0.5 ± 0.1,respectively.

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143140

and l* ¼ 0.25, conjugate faults formed at the velocity discontinuity.For l* ¼ 0.25, one of the conjugate faults formed perpendicular tothe slope. As a consequence, the dihedral angle between the con-jugate faults is smaller than for l* ¼ 0. Mourgues and Cobbold(2003) also made the same observation in their sand models, andexplained this decrease of the dihedral angle as a consequence of anincrease of the internal friction for very low effective stresses (due

Fig. 9. Fault traces observed in the analogue experiments (dashed lines). Fault traces predicthe cohesion of 140 Pa assumed in the analogue tests (bottom panel).

to pore pressure). If we assume that the principal effective stress s1bisects the dihedral angle, we can also notice a small rotation of thestresses (Fig. 8) in response to the seepage forces (Mourgues andCobbold, 2003). For l* ¼ 0 and l* ¼ 0.25, the faults do not seemto be listric, which is not the case for the experiment involving thehighest pore pressure l* ¼ 0.5. In this experiment, the normal faultis clearly listric, very low angle and long.

ted by the analytical model for a hypothetical cohesion of 100 Pa (upper panel), and for

Page 8: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143 141

Based on the analytical model, we determined the fault profilesexpected for the different experimental tests (Fig. 9). Two co-hesions were tested: 100 and 140 Pa. The predictions forCo ¼ 100 Pa seem to be in better agreement with the experimentaldata than for Co ¼ 140 Pa. For l* ¼ 0 and l* ¼ 0.25, the analyticalmodel predicts the formation of normal faults with a very smallcurvature. The dips obtained for Co ¼ 100 Pa are in good agreementwith the experimental faults. Where Co ¼ 140 Pa, the faults seemtoo steep. For l* ¼ 0.5, comparison with the analytical model failedregardless of the cohesion value. By slightly increasing the porepressure, l* exceeds the critical pressure l*c to form a detachment atthe base or within the model, so that the fault can migrate far fromthe velocity discontinuity. With Co ¼ 140 Pa and l* ¼ 0.6, thedetachment occurred at the base of the model. Nevertheless, thepredicted fault is still too steep. For Co ¼ 100 Pa and l* ¼ 0.6,Zc ¼ 3.6 cm, so that a detachment parallel to the surface may form.The predicted shape of the normal fault for these conditions is inrelatively good agreement with the experimental fault. The verylow dip of the experimental fault is probably a good indication ofthe instability of the surface during extension. Nevertheless, adetachment parallel to the surface was not strictly observed,probably because the fault must root at the velocity discontinuity.This velocity discontinuity may also introduce stress perturbationsin the model that could explain some of the discrepancies betweenthe analogue results and the mathematically modelled faults.

5. Conclusions

We used an alternative derivation of the critical Coulombwedgeto analyse the structural effects of cohesion in an overpressuredwedge subjected to gravitational deformation. The stabilizing effectof cohesion is accentuated in the foremost thin domain of thewedge,requiring aMinimum Failure Length (MFL) that controls the locationof the first normal faults to form. This MFL decreases with smallercohesion, smaller coefficient of internal friction, larger fluid over-pressure ratio and steeper upper and basal surfaces of the wedge.Despite the mechanical properties of the wedge being assumed ashomogeneous, non-linearity induced by cohesion results in the for-mation of listric faults. Their location within the wedge and fluidoverpressure ratio control the fault shape. We observe two kinds oflistricity: curvature close to the surface or curvature at depth,depending on cohesion, pore-fluid overpressure, and stability of thesurface. Pronounced curvatures of the faults reflect a high degree ofinstability of surficial part of the wedge. This last prediction wasverified with experiments involving pore pressure, but further ana-lyses and comparisonwith natural examples are required to validateits use as a criterion in structural analysis to evaluate slope stability.

Acknowledgments

ACGC benefited from a PhD scholarship funded by FCT (SFRH/BD/68983/2010). FOM benefited from a sabbatical fellowship(SFRH/BSAB/1405/2014) awarded by FCT, Portugal. The experi-ments and the analytical developments were done at the Universit�eduMaine (France).We thank C�edric Bulois and Christelle Gruber forthe support in the experimental work.

Appendix A. Solution to the force balance including cohesiveand non cohesive materials

We consider a wedge (Fig. 1a) with an upper surface angle a andbasal surface angle b, both surfaces dipping in the same sense, andthus both angles are considered as positive. The Cartesian coordinatesystemx,z is definedwith componentsparallel (x) andperpendicular(z) to the wedge's upper surface. The Cartesian coordinate system

x1,z1 is definedwith components parallel (x1) and perpendicular (z1)to the wedge's basal surface. The prism length (l) and height (h) aredefined perpendicularly to the wedge's basal surface (Fig. 1b).

Following Mourgues et al. (2014), we establish the equations ofstatic equilibrium for the system relative to the Cartesian coordi-nate system x,z, and we define as boundary conditions such that:(1) there is no variation of stresses along the x axis; (2) for z¼ 0 theeffective stress components are null; and (3) the fluid overpressureratio l* is constant within the wedge (the pore-fluid pressure isassumed to increase linearly with z).

In that case, stresses s0zz and s0xz are (Dahlen, 1984):

s0zz ¼�1� l*

�r0gz cosðaÞ (A1)

a0xz ¼ r0gz sinðaÞ (A2)

Considering these effective stress components (Eqs. A1 and A2)and that the wedge is on the verge of failure everywhere (Dahlen,1984), the analysis of the Mohr diagram allows the determination ofs'xx. From geometrical considerations, the s0xx solutions can be foundby expressing the radius (r) of the Mohr circle in two ways (Fig. A1).

1­ In triangle O� B� s0o : r ¼ T þ s0o�sin f (A3)

2­ In triangle s0o � A� s0zz : r ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffis0xz

2 þ s0zz � s0o

2��q(A4)

With s0o the center of the circle and T the tensile strengthexpressed in the linear Coulomb failure criterion as:

T ¼ Co tanðfÞ (A5)

By equalizing these two expressions, a second order equation isfound:

As0o2þ Bs0o þ C ¼ 0 (A6)

with

A ¼ cos2ðfÞ (A7)

B ¼ �2�s0zz þ T sin2ðfÞ

�(A8)

C ¼ s0zz2 þ s0xz

2 � T2 sin2ðfÞ (A9)

Discriminant of equation (A6) is

D ¼ B2 � 4AC (A10)

and solutions are:

s0 ¼ �B±ffiffiffid

p

2A(A11)

s0xx is then deduced from:

s0xx ¼ 2s0o � s0zz (A12)

In the case where the wedge is not cohesive (Co ¼ 0), a simpleexpression of s0xx can be found (Mourgues et al., 2014):

s0xx ¼ s0zz

21±sin f

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1� FS2

p

cos2ðfÞ � 1

!(A13)

Afterwards, it is necessary to transform the stress componentsfrom (x,y) coordinate system to (x1,z1) coordinate system in order toproceed to the stability analysis of the wedge along the basal

Page 9: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143142

detachment:

s0x1x1¼s0xxþs0zz

2þs0xx�s0zz

2cosð�2ða�bÞÞþs0xzsinð�2ða�bÞÞ (A14)

s0z1z1¼s0xxþs0zz

2�s0xx�s0zz

2cosð�2ða�bÞÞþs0xzsinð�2ða�bÞÞ (A15)

s0x1z1 ¼s0xx þ s0zz

2sinð�2ða� bÞÞ þ s0xz cosð�2ða� bÞÞ (A16)

The assessment of the stability state of the wedge is madethrough the balance between the forces driving (Fd) and inhibiting(Fb) the downslopemovement along the basal detachment (Fig.1b).

Fd is the result of the sum of the driving force components alongx1 axis (Fig. 1b):

Fd ¼ Fs þ Fx1w þ Fx1s (A17)

a) Fs corresponds to the integral of s0x1x1 along the local wedgeheight (h, along the z1 axis):

Fs ¼Zh0

s0x1x1dz1 (A18)

b) Fwx1 at a point along the basal detachment of length l and heighth corresponds to the component along �1 of the gravitational loadof the wedge section comprehended between that point and thewedge tip (l¼0, andh¼0). Considering a unitarywidthwedge slice:

Fx1w ¼ 0:5r0ghl sinðbÞ (A19)

c) Fsx1 is the component of the seepage force (SF) parallel to x1,considering a wedge slice of unitary width:

Fs ¼ 0:5l*r0gh2 cosðaÞcosða� bÞ (A20)

in which l* is the fluid overpressure ratio within the wedge.Fb corresponds to the integral of the basal friction on the basal

Fig. A1. Graphical determinatio

detachment:

Fb ¼Z l0

mbs0n dx1 (A21)

inwhich mb and l*b correspond to the internal friction coefficientand the fluid overpressure ratio along the basal detachment,respectively; s'n is the normal effective stress acting on the basaldetachment of the wedge, taking into account a possible differencebetween l* and l*b.

s0n ¼ s0z1z1 � ðPb� PÞ ¼ s0z1z1 þ�l* � l*b

�r0gz cosðaÞ (A22)

where Pb is the pore pressure within the detachment and P thepore-fluid pressure just above the detachment within the wedge.

In case where there is no cohesion, simple expressions of s0x1x1 ,s0z1z1 and s0x1z1 can be found from equations A1-2, A13 and A14-A16:

s0x1x1 ¼ E3r0gz cosðaÞ (A23)

s0z1z1 ¼ E1r0gz cosðaÞ (A24)

s0x1z1 ¼ E2r0gz cosðaÞ (A25)

Fs and Fb are then analytically solved:

Fs ¼ 0:5E3r0gh2 cosðaÞcosða� bÞ (A26)

Fb ¼ 0:5mb�E1 � l* þ l*b

�rgl2 cosðaÞsinða� bÞ (A27)

Finally, the exact solution of the force balance is easily found. Itsexpression is given in the main text (Eq. (17)).

For a cohesive wedge, there is no simple expression of s0xx, s0x1x1or s0z1z1. Consequently, Fd and Fbmust be numerically integrated andthe final balance of forces (Eq. (13)) must be numerically solved.

n of s0xx using Mohr circles.

Page 10: Journal of Structural Geology · Journal of Structural Geology 87 (2016) 134e143 submarine structures at various scales: e.g., large-scale flank col- lapses in volcanic islands (Iverson,

R. Mourgues et al. / Journal of Structural Geology 87 (2016) 134e143 143

References

Bryn, P., Berg, K., Forsberg, C.F., Solheim, A., Kvalstad, T.J., 2005. Explaining theStoregga slide. Mar. Pet. Geol. 22, 11e19. http://dx.doi.org/10.1016/j.marpetgeo.2004.12.003.

Cobbold, P.R., Castro, L., 1999. Fluid pressure and effective stresses in sandboxmodels. Tectonophysics 301, 1e19.

Cobbold, P.R., Durand, S., Mourgues, R., 2001. Sandbox modelling of thrust wedgeswith fluid-assisted detachments. Tectonophysics 334, 245e258.

Cobbold, P.R., Mourgues, R., Boyd, K., 2004. Mechanism of thin-skinned detachmentin the Amazon fan: assessing the importance of fluid overpressure and hy-drocarbon generation. Mar. Pet. Geol. 21, 1013e1025.

Cobbold, P.R., Zanella, A., Rodrigues, N., Loseth, H., 2013. Bedding-parallel fibrousveins (beef and cone-in-cone): worldwide occurrence and possible significancein terms of fluid overpressure, hydrocarbon generation and mineralization. Mar.Pet. Geol. 43, 1e20. http://dx.doi.org/10.1016/j.marpetgeo.2013.01.010.

Costa, A.C.G., Marques, F.O., Hildenbrand, A., Sibrant, A.L.R., Catita, C.M.S., 2014.Large-scale catastrophic flank collapses in a steep volcanic ridge: the PicoeFaialRidge, Azores Triple Junction. J. Volcanol. Geotherm. Res. 272, 111e125. http://dx.doi.org/10.1016/j.jvolgeores.2014.01.002.

Dahlen, F.A., 1984. Noncohesive critical coulomb wedges: an exact solution.J. Geophys. Res. 89 (B12), 10125e10133.

Dahlen, F.A., 1990. Critical taper model of fold-and-thrust belts and accretionarywedges. Annu. Rev. Earth Planet. Sci. 18, 55e99.

Dahlen, F.A., Suppe, J., Davis, D., 1984. Mechanics of fold-and-Thrust belts andaccretionary wedges: cohesive Coulomb theory. J. Geophys. Res. 89 (B12),10087e10101.

Davis, D., Suppe, J., Dahlen, F.A., 1983. Mechanics of fold-and-thrust belts andaccretionary wedges. J. Geophys. Res. 88 (B2), 1153e1172.

Day, S.J., 1996. Hydrothermal pore fluid pressure and the stability of porous,permeable volcanoes. In: McGuire, W.J., Jones, A.P., Neuberg, J. (Eds.), VolcanoInstability on the Earth and Other Planets. Geological Society Special Publica-tion, London, pp. 77e93.

del Potro, R., Hürlimann, M., Pinkerton, H., 2013. Modelling flank instabilities onstratovolcanoes: parameter sensitivity and stability analyses of Teide, Tenerife.J. Volcanol. Geotherm. Res. 256, 50e60. http://dx.doi.org/10.1016/j.jvolgeores.2013.02.003.

Elsworth, D., Voight, B., 1996. Evaluation of volcano flank instability triggered bydyke intrusion. In: McGuire, W.J., Jones, A.P., Neuberg, J. (Eds.), Volcano Insta-bility on the Earth and Other Planets. Geological Society Special Publication,London, pp. 45e53.

Flemings, P.B., Long, H., Dugan, B., Germaine, J., John, C.M., Behrmann, J.H.,Sawyer, D., IODP Expedition 308 Scientists, 2008. Pore pressure penetrometersdocument high overpressure near the seafloor where multiple submarinelandslides have occurred on the continental slope, offshore Louisiana, Gulf ofMexico. Earth Planet. Sci. Lett. 269 (3e4), 309e325. http://dx.doi.org/10.1016/j.epsl.2007.12.005.

Haflidason, H., Sejrup, H.P., Nygard, A., Mienert, J., Bryn, P., Lien, R., Forsberg, C.F.,Berg, K., Masson, D., 2004. The Storegga slide: architecture, geometry and slidedevelopment. Mar. Geol. 213, 201e234. http://dx.doi.org/10.1016/j.margeo.2004.10.007.

Hildenbrand, A., Marques, F.O., Catal~ao, J., Catita, C.M.S., Costa, A.C.G., 2012. Large-scale active slump of the southeastern flank of Pico Island, Azores. Geology 40(10), 939e942. http://dx.doi.org/10.1130/G33303.1.

Hubbert, M.K., Rubey, W.W., 1959. Role of fluid pressure in mechanics of overthrustfaulting. Geol. Soc. Am. Bull. 70, 115e166.

Iverson, R.M., 1995. Can magma-injection and groundwater forces cause massivelandslides on Hawaiian volcanoes? J. Volcanol. Geotherm. Res. 66, 295e308.

Kvalstad, T.J., Andresena, L., Forsberga, C.F., Bergb, K., Brynb, P., Wangen, M., 2005a.The Storegga slide: evaluation of triggering sources and slide mechanics. Mar.Pet. Geol. 22, 245e256.

Kvalstad, T.J., Nadim, F., Kaynia, A.M., Mokkelbost, K.H., Bryn, P., 2005b. Soil con-ditions and slope stability in the Ormen Lange area. Mar. Pet. Geol. 22, 299e310.http://dx.doi.org/10.1016/j.marpetgeo.2004.10.021.

Lacoste, A., Loncke, L., Chanier, F., Bailleul, J., Vendeville, B.C., Mahieux, G., 2009.Morphology and structure of a landslide complex in an active margin setting:the Waitawhiti complex, North Island, New Zealand. Geomorphology 109,184e196. http://dx.doi.org/10.1016/j.geomorph.2009.03.001.

Lacoste, A., Vendeville, B.C., Loncke, L., 2011. Influence of combined incision and

fluid overpressure on slope stability: experimental modelling and natural ap-plications. J. Struct. Geol. 33, 731e742.

Lacoste, A., Vendeville, B.C., Mourgues, R., Loncke, L., Lebacq, M., 2012. Gravitationalinstabilities triggered by fluid overpressure and downslope incisiondInsightsfrom analytical and analogue modelling. J. Struct. Geol. 42, 151e162.

Lehner, F.K., 1986. Comments on “noncohesive critical Coulomb wedges: an exactsolution” by F.A. Dahlen. J. Geophys. Res. 91 (B1), 793e796.

Mandl, G., Crans, W., 1981. Gravitational gliding in deltas. In: McClay, K.R., Price, N.J.(Eds.), Thrust and Nappe Tectonics, Geological Society Special Publication, vol.9, pp. 41e54.

Masson, D.G., Harbitz, C.B., Wynn, R.B., Pedersen, G., Lovholt, F., 2006. Submarinelandslides: processes, triggers and hazard prediction. Philos. Trans. R. Soc. A364, 2009e2039. http://dx.doi.org/10.1098/rsta.2006.1810.

Moore, J.G., Clague, D.A., 2002. Mapping the Nuuanu and Wailau landslides inHawaii. In: Takahashi, E., Lipman, P.W., Garcia, M.O., Naka, J., Aramaki, S. (Eds.),Hawaiian Volcanoes: Deep Underwater Perspectives, Geophysical MonographSeries, 128. American Geophysical Union, Washington D.C., pp. 223e244. http://dx.doi.org/10.1029/GM128

Moore, J.G., Clague, D.A., Holcomb, R.T., Lipman, P.W., Normark, W.R., Torresan, M.E.,1989. Prodigious submarine landslides on the Hawaiian Ridge. J. Geophys. Res.Solid Earth 94 (B12), 17465e17484. http://dx.doi.org/10.1029/JB094iB12p17465.

Morgan, J.K., Moore, G.F., Clague, D.A., 2003. Slope failure and volcanic spreadingalong the submarine south flank of Kilauea volcano, Hawaii. J. Geophys. Res. 108(B9), 2415. http://dx.doi.org/10.1029/2003JB002411.

Mourgues, R., Cobbold, P.R., 2003. Some tectonic consequences of fluid over-pressures and seepage forces as demonstrated by sandbox modeling. Tecto-nophysics 376, 75e97.

Mourgues, R., Cobbold, P.R., 2006a. Sandbox experiments on gravitational spreadingand gliding in the presence of fluid overpressures. J. Struct. Geol. 28, 887e901.

Mourgues, R., Cobbold, P.R., 2006b. Thrust wedges and fluid overpressures: sandboxmodels involving pore fluids. J. Geophys. Res. 111, B05404. http://dx.doi.org/10.1029/2004JB003441.

Mourgues, R., Lecomte, E., Vendeville, B., Raillard, S., 2009. An experimentalinvestigation of gravity-driven shale tectonics in progradational delta. Tecto-nophysics 474, 643e656.

Mourgues, R., Lacoste, A., Garibaldi, C., 2014. The Coulomb critical taper theoryapplied to gravitational instabilities. J. Geophys. Res. Solid Earth 119 (1),754e765. http://dx.doi.org/10.1002/2013JB010359.

Quidelleur, X., Hildenbrand, A., Samper, A., 2008. Causal link between Quaternarypaleoclimatic changes and volcanic islands evolution. Geophys. Res. Lett. 35,L02303. http://dx.doi.org/10.1029/2007GL031849.

Reid, M.E., 2004. Massive collapse of volcano edifices triggered by hydrothermalpressurization. Geology 32 (5), 373e376. http://dx.doi.org/10.1130/G20300.1.

Satake, K., Smith, J.R., Shinozaki, K., 2002. Three-dimensional reconstruction andtsunami model of the Nuuanu and Wailau giant landslides, Hawaii. In:Takahashi, E., Lipman, P.W., Garcia, M.O., Naka, J., Aramaki, S. (Eds.), HawaiianVolcanoes: Deep Underwater Perspectives, Geophysical Monograph Series 128.American Geophysical Union, Washington D.C., pp. 333e346. http://dx.doi.org/10.1029/GM128

Smith, D.E., Harrison, S., Jordan, J.T., 2013. Sea level rise and submarine mass failureon open continental margins. Quat. Sci. Rev. 82, 93e103. http://dx.doi.org/10.1016/j.quascirev.2013.10.012.

Voight, B., Elsworth, D., 1997. Failure of volcano slopes. G�eotechnique 47 (1), 1e31.Walder, J., Nur, A., 1984. Porosity reduction and crustal pore pressure development.

J. Geophys. Res. 89 (B13), 11539e11548.Wang, K., He, J., Hu, Y., 2006. A note on pore fluid pressure ratios in the Coulomb

wedge theory. Geophys. Res. Lett. 33, L19310.Yuan, X.P., Leroy, Y.M., Maillot, R., 2015. Tectonic and gravity extensional collapses in

overpressured cohesive and frictional wedges. J. Geophys. Res. http://dx.doi.org/10.1002/2014JB011612.

Zanella, A., Cobbold, P.R., Rojas, L., 2014a. Beef veins and thrust detachments in EarlyCretaceous source rocks, foothills of Magallanes-Austral Basin, southern Chileand Argentina: structural evidence for fluid overpressure during hydrocarbonmaturation. Mar. Pet. Geol. 55, 250e261. http://dx.doi.org/10.1016/j.marpetgeo.2013.10.006.

Zanella, A., Cobbold, P.R., Le Carlier de Veslud, C., 2014b. Physical modelling ofchemical compaction, overpressure development, hydraulic fracturing andthrust detachments in organic rich source rock. Mar. Pet. Geol. 55, 262e274.http://dx.doi.org/10.1016/j.marpetgeo.2013.12.017.