You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/263807576

Fault-related folding: A review of kinematic models and their application

Article in Earth-Science Reviews · July 2014


DOI: 10.1016/j.earscirev.2014.06.008

CITATIONS READS

153 8,012

2 authors:

Christian Brandes David C. Tanner


Leibniz Universität Hannover Leibniz Institute for Applied Geophysics
94 PUBLICATIONS 1,655 CITATIONS 186 PUBLICATIONS 1,810 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by David C. Tanner on 18 October 2021.

The user has requested enhancement of the downloaded file.


Earth-Science Reviews 138 (2014) 352–370

Contents lists available at ScienceDirect

Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev

Fault-related folding: A review of kinematic models and their application


Christian Brandes a,⁎, David C. Tanner b
a
Institute for Geology, Leibniz Universität Hannover, Callinstr. 30, 30167 Hannover, Germany
b
Leibniz Institute for Applied Geophysics (LIAG), Stilleweg 2, 30655 Hannover, Germany

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

Article history: Folding that is directly related to fault activity is an important deformation feature that occurs all over the world
Received 3 December 2013 in mountain belts, accretionary wedges, fold-and-thrust belts, and intra-plate settings in either strike-slip,
Accepted 27 June 2014 compressional, or extensional regimes. Due to their widespread occurrence, knowledge about the development
Available online 4 July 2014
of these structures is important to a broad spectrum of geoscience sub-disciplines, such as structural geology,
seismology, geomorphology, petroleum geology, and Quaternary geology. Fault-related folding has been
Keywords:
Kinematic model
analyzed intensively over the last 140 years. For the sake of this review, we divide the folds according to the
Fault-related folding way the faults and the folds form; that is into detachment, fault-bend, and fault-propagation folds.
Fault-propagation fold All fault-related folds are caused by changes in fault parameters. The simplest method to produce folds is to trans-
Fault-bend fold port material along faults that have stepped, flat–ramp–flat geometries (fault-bend fold). Alternatively the slip
Detachment fold can decrease along the length of the fault, and depending on whether the fault remains within a detachment
Trishear layer or steps up through mechanical stratigraphy, either a detachment fold or a fault-propagation fold is formed,
respectively.
Detachment folding was first investigated in the early 20th Century, whereas the full significance of fault-
propagation folds was recognized quite late in the 1980s. Seminal work on fault-related folding was carried
out in the 1930s, but quantitative kinematic models have only been available in the last 30 years. These models
are extremely valuable, because they allow a comprehensive understanding of the evolution of fault-related
folds and lead to more accurate predictions of the sub-surface structure. From the mid-1990s onwards, numerical
simulations have been used to identify how fault parameters (such as dip and fault-bend angle, propagation-to-
slip ratio, and shape of the trishear zone) influence the geometry of the related folding. This is directly applicable
to the analysis of the shape of anticlines produced. However, this does not mean that fold geometry is uniquely
related to fault geometry; on the contrary, different kinematic approaches can lead to a similar fold shape.
© 2014 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
1.1. Aims and scope of this work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
1.2. Why is fault-related folding important? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
2. The history of fault-related folding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
2.1. The early years . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
2.2. The period of application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
2.3. Rise of the kinematic approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
3. The definition of geometrical, kinematic and dynamic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 356
4. Kinematic models of fault-related folding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
4.1. Drag folds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
4.1.1. Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
4.1.2. Kinematic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
4.2. Detachment folds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
4.2.1. Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
4.2.2. Kinematic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357

⁎ Corresponding author. Tel.: +49 511 762 4391; fax: +49 511 762 2172.
E-mail address: brandes@geowi.uni-hannover.de (C. Brandes).

http://dx.doi.org/10.1016/j.earscirev.2014.06.008
0012-8252/© 2014 Elsevier B.V. All rights reserved.
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 353

4.3. Fault-bend folds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359


4.3.1. Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359
4.3.2. Kinematic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359
4.4. Fault-propagation folds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
4.4.1. Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
4.4.2. Kinematic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
4.4.3. Velocity models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
4.4.4. Trishear kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
5. Folding mechanisms, distribution of deformation and growth-strata evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
6. Fault-related folding in extensional domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
7. Fault-related folding and the structure of deforming wedges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
8. Fault-related folding and inversion structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
9. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
10. The future of fault-related folding kinematic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
11. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367

1. Introduction different kinematic models, and to show the relevance of fault-related


folding in several geosciences subdisciplines. This review aims to syn-
Numerous outcrops and seismic reflection lines show the intimate thesize the state of the art by presenting the first comprehensive review
association of faults and folds; this is termed fault-related folding of the full range of fault-related folding kinematics. Previous reviews
(Fig. 1). In this work we discuss the geometrical relationship of the have only focused on individual kinematic approaches. We refer to the
two structural elements and show the various kinematic models that simplest kinematic solutions of the models because this makes it easier
have been produced to explain them. Folds that are related to faults to understand. However this also reduces the degrees of freedom and
were recognized early on in the history of geological science (Buxtorf, therefore the necessary variables. In addition, we treat both contrac-
1916; Rich, 1934; Rettger, 1935) and quantitative kinematic models tional and extensional structures, where applicable. Finally we identify
have been available since the mid-80s onwards (e.g. Suppe, 1983; the limitations of the current models and give an outlook on future
Suppe and Medwedeff, 1990; Epard and Groshong, 1995; Poblet and fault-related folding research.
McClay, 1996; Mitra, 2003).
The number of different fault-related fold types depends on how
strictly the group is defined. Since folds have to evolve, not only geomet- 1.2. Why is fault-related folding important?
rically, but also temporally linked to fault movement (as summarized by
Suppe and Medwedeff, 1990), then three kinds of fault-related folds can Fault-related folding is a significant deformation mechanism that
be defined: detachment folds, fault-bend folds, and fault-propagation folds occurs in mountain belts and foothills (e.g., Boyer, 1986; Vann et al.,
(Jamison, 1987). Hybrid structures are also possible (Marrett and 1986; Erslev and Mayborn, 1997; Ford et al., 1997; Homza and
Bentham, 1997). Drag folding can be added to this list because it is also Wallace, 1997; Delcaillau et al., 1998; Burbank et al., 1999; Tozer et al.,
temporally and geometrically-related to faulting. 2002; McClay, 2011; Calamita et al., 2012) (Fig. 1A, B), deepwater
Fault-related folds are, in general, due to the changes in fault param- fold-and-thrust belts (Corredor et al., 2005; Morley et al., 2011), accre-
eters. Thus the simplest method to produce folds is to transport material tionary wedges (Biju-Duval et al., 1982; Gulick et al., 2004), intraplate
along faults that have stepped, flat–ramp–flat geometries (fault-bend settings (Bump, 2003), and during basin inversion (Mitra, 1993;
fold model; Fig. 2C). Alternatively the slip can decrease along the fault's Okamura et al., 2007). The beauty of the concept of fault-related folding
length, which is either compensated by folding at the fault tip, if the is that it can explain the evolution of folds in the upper crustal levels,
fault cuts through stratigraphy (fault-propagation fold model; Fig. 2D), where brittle processes are typically expected to exist (Suppe, 1983).
or if instead the fault does not ramp up, a detachment fold may form Besides its dominance in compressional regimes, it is also an important
(Fig. 2B). deformation process in extensional settings (Howard and John, 1997;
Secondary faults that form as the result of strain discontinuities Sharp et al., 2000; Ferrill et al., 2012) (Fig. 1C, D), where it is often
during folding are named fold-accommodation faults (Mitra, 2002a). expressed either as roll-over anticlines on listric normal faults (Song &
They usually show distinctly smaller amounts of slip than the fold- Cawood, 2001) (Fig. 1E), as extensional fault-propagation folds, or as
forming faults. Fold-accommodation faults form as a consequence of extensional forced folds.
increased curvature of folds, shear of forelimbs, or as back-thrusts In addition, fault-related folding can occur in strike-slip situations
that accommodate hanging-wall strain during fault-related folding (e.g., Tindall and Davis, 1999). It is also significant in the field of deep-
(Mitra, 2002a). In general, fault-related folds can form on a kilometer- water fold-and-thrust belts (Camerlo and Benson, 2006; Morley et al.,
(Apotria and Wilkerson, 2002; Korsch, 2004) to meter-scale 2011), soft-sediment deformation, gravity-driven slumps, and mass-
(McConnell et al., 1997). transport complexes (Alsop and Marco, 2012), as well as in ice-sheet
induced glaciotectonics (Thomas and Chiverrell, 2007; Burke et al.,
2009; Brandes and Le Heron, 2010). It has been well known for a long
1.1. Aims and scope of this work time that anticlines represent traps for oil and gas (Hopkins, 1917;
Rich, 1921). Therefore fault-related folding plays a great role in the for-
Over the last 30 years, fault-related folding has been treated in a mation of structural hydrocarbon reservoirs (Mitra, 1990; Ingram et al.,
kinematic context. The aim of this paper is to summarize and review 2004; Kent and Dasgupta, 2004). It also acts as an important co-seismic
these models. We will focus on the kinematic evolution and the deformation process (e.g. Kao, 2000; Chen et al., 2001; Johnson and
resulting geometry of fold structures. The intention is to stress the role Segall, 2004; Guzofski et al., 2007; Lin et al., 2007). Consequently,
of fault-related folding as one of the most important deformation fault-related folding is of great economic interest and represents a cru-
mechanisms that globally occur in various tectonic settings, to review cial parameter in the evaluation of earthquake hazards (Allmendinger
354 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

TWT
A) B)

fault
trace

fault tr ace 2

C) D)

fault tr ace

E)

roll-over anticline

fault tr ace

Fig. 1. Different types of fault-related folds. A) Fold in the hanging wall of a large thrust sheet, Alps, southern Germany. B) Seismic section showing a fault-propagation fold, Limón fold-and-
thrust belt, offshore southeastern Costa Rica. Modified after Brandes et al. (2007a). C) and D) Extensional fault-propagation fold, northern foreland basin of the Alps, southern Germany.
E) Syn-tectonic growth fault in Namurian siliclastic sediments exposed on the west coast of Ireland, Donegal Point, County Clare.

and Shaw, 2000; DeVecchio et al., 2012). Fault-related folding also influ- petroleum geology and Quaternary geology. Kinematic models (see
ences the evolution of the Earth's surface (Jackson et al., 1996; Mueller Table 1), of fault-related folding can be utilized in all fields of applied ge-
and Suppe, 1997; Delcaillau et al., 2006) by its impact on the evolution ology, in which an understanding of the deformation in the subsurface is
of the drainage pattern (Keller et al., 1998), and it would even seem that required (such as hydrocarbon/geothermal energy exploration and CO2
it affects the topography on other planets (Okubo and Schultz, 2004). sequestration). The limits in visualization deformation in the vicinity of
This shows that not only is fault-related folding relevant to the field of faults on seismic surveys due to the so-called “fault-shadow” the correct
pure structural geology, but it is also highly pertinent to a wide range citation is (Aikulola et al., 2010) can be overcome by kinematic models.
of geoscience subdisciplines, such as seismology, geomorphology, Kinematic models allow to forward model the position of the faults,
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 355

A) undeformed state Fischer & Wilkerson, 2000), but it has been also shown that these sym-
metric fracture-fold models can fail (Reches, 1976). Kinematic fault-
related fold models can relate the anticline shape to the fault geometry
and therefore serve as a tool for fracture prediction.

2. The history of fault-related folding

2.1. The early years

B) detachment fold The fact that faults and folds could be linked was recognized early on
in geological research (Fig. 3). Willis (1894) introduced the break-thrust
concept, in which folding and faulting are directly related. He concluded
that the rocks were initially folded, and, if the deformation continued,
the shortening was compensated by faulting. This is not strictly fault-
related folding because in Willis' model, faulting follows folding.
Cadell (1889) presented analogue models where faulting and folding
detachment could be observed as closely-linked processes. Later, Chamberlin
(1910) introduced the first method to calculate the position of a basal
detachment, based on the geometry of folds exposed at the Earth's
C) fault-bend fold
surface. Buxtorf (1916) showed an interpretation of the Swiss Jura
Mountains, with several detachment folds, highlighting the regional
importance of this type of deformation. Stille (1924) defined the so-
called german-type paratectonics (= type of intraplate deformation in
northern Germany), where faults and folds are the dominating structur-
al elements. Benchmarking work on fault-related folding was presented
fault bend by Rich (1934). He analyzed folds in the Appalachians and concluded
that anticlines in a fold belt were the result of thrust sheets moving
D) fault-propagation fold over ramps. This model is still valid and the resulting structures are
known today as fault-bend folds. Direct proof that faulting can cause
folding was presented by Rettger (1935), based on analogue models.
Cloos (1936) presented generic models to show that folding can
be the consequence of faulting (Fig. 4). He interpreted folding in the
hanging-walls and footwalls of large thrust faults to be related to drag.
The term Bruchfaltengebirge (lit. — “brittlely-folded mountain belt”)
fault tip was explicitly used for areas in which folds are related to faults (Cloos,
1936).
Fig. 2. The three main types of fault-related folds. A) Undeformed state. B) A detachment
fold evolve from the shortening of a rock mass above a detachment. The space in the fold
core is filled with mobile material. C) A fault-bend fold forms when material is transported 2.2. The period of application
over a bend in a fault. This forces the material into an antiformal geometry. D) Fault-
propagation folds form as a consequence of variation in the slip along the fault. The slip The 1960–80s were characterized by strong advances in the quanti-
on a fault is assumed to decrease to zero at the tip, which is compensated by buckling of fication of deformation processes. Several studies dealt with structural
material above the fault.
restoration techniques (Dahlstrom, 1969; Hossack, 1979; Elliott, 1983)
and fault-related folds played a major role in quantifying shortening
based on the geometry of the hanging-wall deformation. Another appli- in fold-and-thrust belts (Cooper and Trayner, 1986; DePaor, 1988;
cation for kinematic models of fault-related folding is in the field of pe- Dahlstrom, 1990; Wu et al., 2005).
troleum geology for fracture prediction. In many cases, hydrocarbons
are trapped in fractured reservoirs and fractures are important migra- 2.3. Rise of the kinematic approach
tion pathways (e.g. Aydin, 2000; Nelson, 2001), but in general, such
fractures are below the resolution of standard geophysical methods From the early 1980s onwards, fault-related folding began to be
and their orientation and extent is rather difficult to map in the subsur- treated in a kinematic fashion (Fig. 3). Suppe (1983) introduced the
face and can only be predicted by models. The geometry of an anticline first kinematic model of fault-bend folds, based on the concept of
has been put forward as a proxy for fracture distribution (Lisle, 1994; kink-band migration. Later in the 1990s, different kinematic models

Table 1
Glossary of specific expressions used in the text.

• Kinematics: The study of the motion of points within a body without consideration of the cause of the motion.
• Velocity: A vector that is defined by magnitude and direction. It represents the rate of change of the position of an object. Velocity is an important concept in kinematics.
• Flexural-slip: The concept of allowing differential movement to take place on a detachment or multiple detachments.
• Detachment: A discontinuity within a body, usually along planes of mechanical weakness, i.e. bedding, across which points move with different velocities.
• In-plane motion: All the motion that is caused by the deformation is within a two-dimensional plane. This notion is synonymous with bulk plane strain.
• Line-length, area, and volume preservation. This is an attempt to constrain a kinematic algorithm, and therefore make it more realistic, by maintaining line-length and/or area in two-
dimensions or up to all three values in three-dimensions, between, for instance, pins.
• Pins. Virtual lines or planes that are used to divide and bind a model. Fixed pins are set at the boundaries between deformed und undeformed domains or along the axial planes of
folds, while loose pins are allowed anywhere in the model, and may move and possibly rotate with the deformation. Analysis of the pins allows the quantification of deformation as
well as the possibility to judge, e.g. the amount of strain within the model. Subdividing a model with a number of pins allows incremental deformation to be determined.
356 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

extensional fault- trishear kinematics in 3D


propagation folding by Cristallini et al.
by Hardy & McClay
double-edge fault-
Stille´s f ault Suppe & Medwedeff´s propagation folding
and fold dominated kink band migration model by Tavani et al.
deformation style for fault-propagation folds trishear velocity
Chamberlin calculated folding and fracturing description by
detachment depth recognition of trishear kinematics Zehnder & trishear kinematics
of rocks by Ramsay
fault-bend folds by Erslev Allmendinger in 3D by Cardozo
by Rich
analogue models Dahlstrom´s balanced kinematic model
of faults and folds cross-sections for detachment f olds trishear for
analogue models
numerical modelling continuously
by Cadell of fault-related folds by Epard & Groshong
of deformation by curved faults by
by Rettger and velocity description for
Waltham Brandenburg
simple and pure shear by
Willis introduced Buxtorf´s Cloos described Suppe´s kink band Waltham & Hardy
break-thrust Swiss Jura drag folds as fault migration model for
concept section related fault-bend folds

1880 1890 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 2010 2020

Fig. 3. History of research on fault-related folds and their basic kinematic models.

for detachment folds (Epard and Groshong, 1995; Homza and Wallace, significant for the geometric evolution of the related fold (Hardy and
1995; Poblet and McClay, 1996) and fault-propagation folds (Mitra, Ford, 1997; Allmendinger, 1998). The propagation-to-slip ratio was
1990; Suppe and Medwedeff, 1990; Erslev, 1991; Mitra and Mount, shown by Chapman and Williams (1984), using displacement–distance
1998) were developed. These 2D models were then extended to pseudo graphs, to be a key parameter of fault-propagation folds. Displacement–
3D (Wilkerson et al., 1991). At the same time, contractional fault-related distance plots have been used also to compare fault-related folding pro-
folds were successfully analyzed using analogue sandbox models cesses that occur in natural structures observed at different scales
(McClay, 1995; Storti et al., 1997; Bernard et al., 2007). The use of (Tavarnelli, 1997). Also in the late 1990s, fault-related folding in exten-
numerical simulations since the late 1990s has greatly extended knowl- sional regimes moved in the research focus (Howard and John, 1997;
edge on fault-related folds (Hardy and Ford, 1997; Allmendinger, 1998; Hardy and McClay, 1999; Sharp et al., 2000) and attracted interest up
Tanner et al., 2003; Allmendinger et al., 2004; Cardozo and Aanonsen, to the mid-2000s (Khalil and McClay, 2002; Finch et al., 2004; Jin and
2009; Cardozo et al., 2011). Computer simulations led to a more quan- Groshong, 2006). The work of Medwedeff (1989), Suppe et al. (1992,
titative understanding of these structures. Parameters such as the 1997) and Shaw and Suppe (1994, 1996) opened the door for the analy-
propagation-to-slip ratio of the fault were recognized as highly sis of fault-related folds based on their growth strata. In the following
years, growth strata studies increased (Hardy et al., 1996; Ford et al.,
1997; Poblet et al., 1997; Storti and Poblet, 1997; Vergés et al., 2002;
Brandes et al., 2007b). Recent developments in this field are also the
analysis of growth-strata distribution on the evolution of fault-related
folds (Strayer et al., 2004).

3. The definition of geometrical, kinematic and dynamic models

Our model definition follows the criteria of Greenwood (1989), in


that they 1) are based on a guiding conceptual model, 2) have a
theoretically-based submodel, 3) have a predictive capability, and
4) have the ability to deliver testable predictions. Nevertheless it is
entirely possible for different models to produce the same structure. If
a model is unique, then only that model can explain an observation,
whereas if a model is consistent, more than one model can explain the
observation (Stüwe, 2000).
Models that contain a primary dataset (2D or 3D) of stratigraphic
units, faults, and other geological information, as e.g., shown in
Culshaw (2005), can be termed geometrical. Depending on the quality
and extent of a dataset, either a two- or three-dimensional geometrical
model can be built. The accuracy of the dataset, for instance the resolu-
tion of the seismic data used to generate the geometrical model, deter-
mines the final resolution. Similarly, outcrop data can be included in
the model, but they are increasingly limited in their ability to predict
structure at greater depths.
Kinematic models determine the evolution of a geometrical struc-
tural model (e.g. Laubscher, 1985), so that the paths of material points
in the model can be followed and examined. Kinematic indicators,
i.e. information about the kinematic vectors involved in the structural
development, are, for instance, slickensides on faults, axes of folds, cur-
Fig. 4. Drawing by Cloos (1936) that illustrates the evolution of a fold associated with a
vature of faults, and stratigraphic contacts. Kinematic models can be
fault. The picture shows that the fold grows with increasing slip. This early illustration of compounded to include different geological events (i.e. unconformities)
fault-related folding shows faulting and folding as a contemporaneous process. and various kinematic movements (i.e. basin inversion).
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 357

Dynamic models study the forces and stresses required to produce (Fig. 6A). A detachment is defined as a low-angle fault that may be
the structure. They require geometrical and kinematic models as nearly, but not exactly parallel to a horizon, whereas as a décollement
input, as well as rheological information (e.g. density, thermal proper- is a bedding- or layer-parallel fault (Peacock et al., 2000). Detachment
ties, Young's Modulus, and Poisson's ratio) about the rocks involved folds also develop above the tip of a bedding-parallel thrust (Poblet
(e.g. Braun and Sambridge, 1994). Boundary conditions must include and McClay, 1996) in competent rocks and are often cored by less-
gravitational forces and the far-field stress vectors, including paleostress competent rocks (Homza and Wallace, 1995). Distinct detachment
vectors when considering structural development over geological time. folds can only form if there is enough mobile material present in the
detachment layer to fill the core of the growing anticline (Stewart,
4. Kinematic models of fault-related folding 1996). If there is not enough material to fill the amplifying fold, fold
growth will stop and shortening may be compensated by faulting
4.1. Drag folds (Stewart, 1996).

4.1.1. Characteristics 4.2.1. Characteristics


Ramberg (1963) defined drag folds as quasi-monoclinic folds The most important characteristic feature of detachment folds is the
that often develop in thin competent rock layers bounded by more detachment horizon (i.e. décollement) (Fig. 6A) that is developed either
competent thick layers. In this model, drag folds form when a package above, below, or above and below the fold. Layer-parallel strain allows
of alternating competent and less competent layers bends as the result the fold to develop above a fixed stratigraphic detachment (Groshong
of horizontal shortening. More commonly the term drag fold is and Epard, 1994). Detachment folds are often symmetric (Hardy and
used for a fold that developed as a consequence of friction along a fault Finch, 2005) and their geometries range from concentric to chevron-
(e.g. Cloos, 1936), although this is viewed skeptically by Grasemann type folds or even box folds (Rowan et al., 2004; Shaw et al., 2005).
et al. (2005), who propose that drag folds are in fact caused by the Internally, detachment folds are characterized by homogeneous strain,
heterogeneous displacement field around a fault as it undergoes slip. second-order folds, second-order conjugate faults, or duplex structures
Both reverse and normal drag can be distinguished (Fig. 5). Normal (Epard and Groshong, 1995).
drag means that the deformed rock layers are convex in the direction Detachment folds can be subdivided into two main groups: a) dis-
of transport along the fault, whereas in the case of reverse drag they harmonic folds and b) lift-off folds (Mitra and Namson, 1989). Dishar-
are concave in the transport direction (Hamblin, 1965; Grasemann monic folds are parallel or concentric folds with disharmonic folding
et al., 2005). in the core (Fig. 6B). Lift-off folds are parallel folds, where the core of
the anticline is characterized by isoclinal limbs (Fig. 6B). Early models
4.1.2. Kinematic models of detachment folds were geometrical (Jamison, 1987) that focused on
The sense of shear along the fault can be determined by the asym- the shape of the folds. Different kinematic models were developed
metry of the drag fold. Furthermore, the slip vector can be estimated later. From an area-balanced cross-section it is possible to derive the
using Hansen's Method (Hansen, 1971; Twiss and Moores, 1992). If depth of the detachment (Fig. 6C).
the fold is conical, then the fold axis is the axis of rotation of the beds
at the fault, which is perpendicular to the slip direction (Becker, 4.2.2. Kinematic models
1995). Recently, numerical (finite element) models have been used to Epard and Groshong (1995) introduced the limb rotation model,
test drag folding geometries and processes (e.g. Reches and Eidelman, which describes the evolution of detachment folds above a fixed
1995; Resor and Pollard, 2012). These models prove that edge disloca- layer-parallel detachment (see also Hardy and Poblet, 1994). The basic
tion in an elastic half space is sufficient to produce the majority drag assumption of this model is that the deformation takes place between
folds, and that friction on the fault is not important. two pin lines and the displacement along the detachment is compensated
by folding, which means that the model is locally line-balanced. Deforma-
4.2. Detachment folds tion occurs by amplification of the anticline. The model emphasizes
symmetric folds (Epard and Groshong, 1995). The fold in the model has
Detachment folds are folds that evolve from the shortening of a rock three axial planes, a vertical one in the center (corresponding to the
mass above either a detachment or a décollement (e.g. Poblet et al., plane of symmetry) and two vertical or inclined ones that separate
1997; Rowan, 1997; Scharer et al., 2004; Poblet and Hardy, 1995) the fold from the undeformed area (Fig. 6D). Following Epard and

normal fault reverse fault


normal drag
reverse drag

Fig. 5. Drag folds develop along normal faults and reverse faults. They can be classified into drag folds with normal drag and drag folds with reverse drag. Normal drag means that the beds
are convexly warped in the transport direction of the fault while reverse drag means the opposite. Drawn after Grasemann et al. (2005).
358 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

general characteristics of detachment folds B) basic types of detachment folds


A)
disharmonic detachment fold
competent rocks

less competent
rocks

slip
fault tip
detachment

detachment

C) l0 balanced detachment fold


lift-off detachment f old

l0 - l 1 l1

A z
depth to detachment: z = A
l0 - l 1
detachment
detachment

D)
model of Epard & Groshong (1995)

fixed hinge migrating hinge


W
fixed pin line
reference level B B
C A C A

D H D H
s
s h h
Θ Θ

E)
model of Poblet & McClay (1996)

1) constant limb dip/v ariable limb length 2) variable limb dip/constant limb length

3) variable limb dip/v ariable limb length 4) variable limb dip/v ariable limb length

Fig. 6. Detachment folds. A) General characteristics of detachment folds are a horizontal detachment and a core of less-competent rocks. B) Types of detachment folds: 1. fold core
is characterized by disharmonic folds, 2. Lift-off folds, where the core is characterized by isoclinal limbs. C) Depth-to-detachment calculation of a detachment fold, using the
excess-area-method. D) Kinematic model of Epard and Groshong (1995) for fixed and migrating hinges (see text for explanation). E) Kinematic model of Poblet and McClay
(1996), see text for explanation.
Panel A redrawn after Poblet and McClay (1996). Panel C modified after Mitra and Namson (1989), Epard and Groshong (1993) and Wiltschko and Groshong (2012).
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 359

Groshong (1995), the key parameters of the model are the width of theoretical models (Biot, 1961), natural observation (e.g. Daëron et al.,
the fold (W), the dip of the external axial planes (θ), the distance from 2007), and retro-deformation of cross-sections (Tanner et al., 2011).
the reference level to the detachment (H), and the displacement along
the detachment (D). The fold grows by an increase of the area s that is 4.3. Fault-bend folds
equal to Dh (Fig. 6D). Their model describes the evolution of a detach-
ment fold with a fixed hinge, as well as the evolution of a fold with Fault-bend folds occur when material is transported over a thrust
migrating hinge. In case of a fixed hinge, no material crosses the axial footwall ramp (Berger and Johnson, 1980), i.e. a steepening of the
surfaces. The fold growth is the result of limb rotation and changes in fault plane that typically crosscuts stratigraphy rather than following
bed length (Hardy and Poblet, 1994; Epard and Groshong, 1995). the bedding plane (Suppe, 1983) (Fig. 2C). This concept was originally
Homza and Wallace (1995) presented another geometrical and introduced by Rich (1934) based on his observations in the Appala-
kinematic model for the evolution of detachment fold. Their model chians. Wiltschko (1979) developed a mechanical model for a thrust
deals with two different scenarios: a) a detachment fold that develops sheet that moves over a ramp. Fault-bend faults have been recognized
above a detachment unit with constant thickness and b) a detachment in fold-and-thrust belts all over the world and have been the focus of
fold that forms above a detachment unit with changing thickness during many studies (Suppe, 1983; Zoetemeijer et al., 1992; Medwedeff and
deformation. The model of Homza and Wallace (1995) assumes the Suppe, 1997; Savage and Cooke, 2003; Suppe et al., 2004).
conservation of bed length and area. The competent layer produces par- A special type of fault-bend fold is a triangle zone or antiformal
allel folds by flexural-slip, therefore conserving bed length. The constant stack. Triangle zones are common structural elements that occur
detachment depth model produces folds that are initially symmetric globally and are characteristic of the transition between fold-and-
and grow with a fixed arc-length. The variable detachment fold model thrust belts and foreland basins. The term “triangle zone” has been
allows both fold growth with fixed and migrating hinges. given to different types of structures with a triangular shape in the
A third model for detachment folds was presented by Poblet and cross section (e.g., McClay, 1992). In most cases, it describes an
McClay (1996). These authors analyzed three individual kinematic antiformal duplex on the leading edge of a fold and thrust belt that
mechanisms for: a) constant limb dip, b) constant limb length and internally consists of vertically-stacked horses. Different studies have
c) variable limb dip and limb length (Fig. 6E). Constant limb dip focused on the mechanical and kinematic evolution of triangle zones
means that the fold grows by lengthening of the limbs, whereas in (e.g., Erickson, 1995; Erickson and Jamison, 1995; Jamison, 1996;
the constant limb-length mechanism, fold growth is achieved by Jones, 1996; Stockmal et al., 2001; Tanner et al., 2010). Triangle zones
limb rotation. The variable limb length/limb dip mechanism pro- begin as fault-bend folds. After a first fault-bend fold is formed by mate-
duces a detachment fold that grows by limb lengthening and limb ro- rial that is transported over a footwall ramp, the ramp shifts slightly for-
tation. Poblet and McClay (1996) consider four different end- ward, which leads to the vertical stacking of overlapping fault-bend
members (Fig. 6E): No. 1 assumes constant limb dip and variable folds that form the core of the triangle zone (Jones, 1996). An antiformal
limb length. In this case, the dip of the limb and the axial surfaces stack forms when the displacement on each horse equals the ramp
are fixed. The shortening is accommodated by an outward migration spacing (Doglioni and Carminati, 2008).
of the fold boundaries. No. 2 has a constant limb length and fold
growth is caused by steepening of the limb dip (limb rotation). The 4.3.1. Characteristics
stratigraphic thickness is maintained by the rotation of the axial sur- The basic requirements for the evolution of fault-bend folds are
faces. Nos. 3 and 4 are both characterized by variable limb dip and changes in the geometry of the fault plane, such as bends or ramps. A
length. In these latter models, shortening is compensated by limb ro- footwall ramp usually connects two horizontal detachments (flats) at
tation and outward propagation of the axial surfaces. The fold grows different stratigraphic levels (Fig. 7). By moving from the first, deepest
by lengthening of the limbs and steepening of the limb dip. No. 4 is footwall flat to the shallowest, the hanging-wall material will be
similar to No. 3, but in this case, the detachment depth is given by deformed into a syncline–anticline–syncline structure. Another require-
the intersection of the opposing axial surfaces. Poblet and Hardy ment for fault-bend folding is that the fault blocks are in tight contact
(1995) showed that the key aspects of detachment folds, such as along the plane and the bend in the fault does not lead to the develop-
fold amplification and slip along the fault, can be derived from the ment of significant voids (Suppe, 1983). In the case of fault-bend folds,
analysis of the growth-strata geometry that records the rotation of the propagation rate of the fault must be much higher than fault slip,
the anticline limbs. because the fold forms before the first increment of slip (McNaught
Mitra (2002b, 2003) presented a kinematic model for detachment and Mitra, 1993). Bending of the material during the transport over
folds that explains why a large number of geometrically-different the ramp is accommodated by bedding-parallel simple shear and the
detachment folds can occur. Like all previous models, it assumes that magnitude of shear strains is related to the ramp angle (Sanderson,
the fold develops in a competent rock layer that is underlain by a less 1982). This causes beds to thin over a concave fault bend. To produce
competent unit. The fold mainly deforms by flexural slip, with some ad- a fault-bend fault, enough energy must be available to overcome
ditional faulting and fracturing. The wavelength of the fold is controlled the friction on the fault, uplift the fold, and shear the fold material
by the thickness of the involved rock units. Continued deformation (Williams, 1987).
leads to increasing amplitude and wavelength of the fold. This is driven
by the rotation of limb segments and hinge migration (Mitra, 2003). 4.3.2. Kinematic models
The initial limb rotation is a flexural-slip process. Ongoing shortening Suppe (1983) introduced the first, and still one of the most compre-
causes internal deformation between locked hinges. Variations in the hensive kinematic models for fault-bend folds, which is now often
structural style of detachment folds are controlled by the mechanical called the “kink-band migration model”. Folds evolve above bends in
stratigraphy, the amount of shortening, the asymmetry, and the occur- the fault plane (i.e. the bends between the flats and ramps). The basic
rence of faulting (Mitra, 2003). assumption is that the fault consists of straight segments, which are sep-
Contreras (2010) considered detachment folding on the basis of the arated by an angular bend and only the hanging wall block deforms by
conservation of mass, and not on a specific folding mechanism. There- flexural slip; the footwall block remains rigid (Suppe, 1983; Medwedeff
fore in his model fold limbs are not necessarily straight. The model and Suppe, 1997). As the movement begins along the fault, two kink
can be extended to include erosion and sedimentation during the fold- bands evolve, one at the base of the ramp and one at the upper end
ing. From the resulting velocity models, Contreras (2010) has shown (Suppe, 1983) (Fig. 7). The axial surfaces of the kink band at the upper
that detachment fold amplitude increases exponentially over time ramp end are named A and A′. The axial surfaces of the kink band at
under constant shortening, which has been hypothesized from the base of the ramp are named B and B′ (Fig. 7). The axial surface B
360 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

A)

flat ramp flat


B) B A A´

X X´


Y

B´ A
A1 + A2 = A0
γ2 γ
A1 1

C) B A´

ϕ
A2
β

A0 Θ X X´

Fig. 7. Kink-band migration model for a fault bend fold, based on Suppe (1983). Material is transported over a footwall ramp and deformed into an antiform, see text for explanation.

bisects the angle between the flat and the ramp. The axial surfaces A and of cross-sections that can be used for the exploration of structural
B end at the fault bends (points X and Y, respectively). Further move- hydrocarbon traps and the assessment to seismic risks by identifying,
ment causes one of each of the axial surfaces of these folds to migrate for instance, the position of a potentially-seismogenic blind thrust.
in the slip direction, so that the two folds increase in amplitude with Many earthquakes occur along thrusts that are blind, i.e. they lack a
continued slip. The axial surfaces A′ and B′ end at the points X′ and Y′, surface fault trace (Lettis et al., 1997). In these cases near-surface short-
respectively. ening is accommodated by folding (Dolan et al., 2003). There are several
Due to the formation of the upper kink band, the slip along the fault well-studied examples of seismogenic blind thrusts from the Los
decreases on the forelimb side by about 60% (Suppe, 1983). The basic Angeles area (Shaw and Shearer, 1999). For instance, the 1994
pre-assumption for this kinematic model is layer-parallel slip that Northridge earthquake in Southern California took place along a blind
produces parallel folds. This means that bed thickness and length are thrust (Yeats and Huftile, 1995) and the 1987 Whittier Narrows earth-
conserved. The great advantage of the model of Suppe (1983) is that it quake was related to a pronounced, fault-related hanging-wall anticline
relates the geometry of the fold directly to the geometry of the fault. that developed above a blind ramp (Shaw et al., 2002). Namson and
This allows predictions of the subsurface geometry of tectonic struc- Davis (1988) showed that the Coalinga anticline and the Kettleman
tures to be based only on a small set of observations. The model assumes Hills are also fault-related folds that evolved above seismically-active
the following: a) sharp bends in the fault plane up to 30°, b) conserva- faults. Analyzing blind thrusts with standard geological methods,
tion of area and c) constant layer thickness of the beds. This implies such as mapping, is difficult because the faults do not intersect with
conservation of bed length (in the slip direction), layer-parallel slip the surface (Roering et al., 1997), and therefore other approaches have
and the formation of kink folds that have straight limbs and show infi- to be used. For instance, balanced cross-sections make it possible
nite curvature (Suppe, 1983). In such a case, the axial surface of the to identify unknown faults. This was shown by Davis and Namson
fold bisects the angle between the fold limbs. The model then predicts (1994), who constructed a cross-section based on the kink-band meth-
the fault bend angle ϕ, the cut-off angle θ, and the opening angle of od that revealed the seismogenic fault that caused the 1994 Northridge
the fold, characterized by the angle γ: earthquake.
  A modification of the kink-band migration approach is the concept
−1 sin2γ of shear fault-bend folds (Suppe et al., 2004), which considers the inter-
ϕ ¼ θ ¼ tan :
1 þ 2 cos2 γ nal shear that commonly affects thrust sheets. This kinematic model ex-
plains the observation that backlimb angles dip less than the fault ramp
The maximum ramp angle of 30° consequently limits the forelimb does. This problem has been tackled with the concept of shear fault-
angle of the resultant fault-bend fault. The relationship of the fold bend folding, which is a combination of limb rotation and kink-band mi-
shape to the fault-bend angle means that the fault geometry can be pre- gration (Suppe et al., 2004). Two end-member models for shear fault-
dicted from the form of the fold. This is a key method for reconstructing bend folding have been developed (Fig. 8) (Suppe et al., 2004; Shaw
the subsurface geology based on outcrop data. It allows the construction et al., 2005). In the simple-shear case, the décollement layer experiences
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 361

A) Medwedeff and Suppe (1997) presented solutions for fault-bend


simple-shear fault-bend fold
folds, in which more than one bend is developed. Whereas single
bends produce folds that have kink-like shapes, multi-bend faults
will lead to more curved fold hinges. (Medwedeff and Suppe, 1997). A
modified model for fault-bend fold with rounded hinges was developed
by Tavani et al. (2005) using circular hinge sectors, in which the hinge
sectors are pinned to the fault plane and replace the straight axial sur-
décollement
faces of the kink-band migration model. This is a step towards rounded
layer
“natural” fold shapes.
Tavani and Storti (2006) pointed out that the standard kinematic
model for fault-bend folding is limited because it requires the entire
fault to have propagated before any hanging-wall deformation can
begin. They propose that progressive fault propagation is more realistic.
no flat developed ramp
This means that fault-bend folding can be treated as a type of fault-
propagation folding (Tavani and Storti, 2006). With the concept of
B) pure-shear fault-bend fold double-edge fault-propagation folding (Tavani et al., 2006) (for expla-
nation see chapter fault-propagation folding) it is possible to create
realistic fault-bend folds. With small slip-to-propagation ratios a fault-
propagation fold is produced, which when transported forward along
the upper flat, is geometrically similar to standard kink-band migration
fault-bend folds (Tavani and Storti, 2006).
décollement Another kinematic algorithm to simulate fault-bend folding, known
layer as “fault-parallel flow”, was developed by Egan et al. (1997). This
algorithm quite simply states that all material objects within the hang-
ing wall move parallel to the fault surface, along virtual flow paths
(Fig. 10) (Tanner et al., 2003; Ziesch et al., 2014). This is theoretically
applicable to a large range of complex fault geometries, such as rounded
flat ramp fault bends and multiple-dipping fault domains. However, Ziesch et al.
(2014) have shown that the algorithm gives unnaturally high strain in
Fig. 8. Shear fault-bend folding. Two end-member models for shear fault-bend folding have
the hanging walls of extensional faults and is therefore only suitable
been developed. In the simple-shear case, the décollement layer experiences bedding-
parallel shear, in the pure-shear case, the décollement layer moves above a flat, as in the for fault bend angles of less than 40°, i.e. typically compressional
no-shear fault-bend fold model, but instead shortens and thickens above the ramp. structures.
Modified after Suppe et al. (1992) and Shaw et al. (2005).
4.4. Fault-propagation folds
bedding-parallel shear (Fig. 8A) (Suppe et al., 2004; Shaw et al., 2005).
In the pure-shear case, the décollement layer moves above a flat, as in Fault-propagation folds may also form in the hanging walls of faults
the non-shear fault-bend fold model, but also shortens and thickens (Fig. 11). However these folds are different from fault-bend folds,
above the ramp (Fig. 8B) (Shaw et al., 2005). By unfolding the hanging because they form as a consequence of variations in the slip along the
wall, the resulting shearing of the back pin gives the shear profile. fault. The slip on a fault is assumed to decrease to zero at the tip. This
From such a shear profile it is possible to derive the amount and position decrease in slip is compensated by buckling of material above the fault
of shear in the deformed package (Suppe et al., 2004). (Williams and Chapman, 1983; Suppe and Medwedeff, 1990). In con-
A kinematic velocity field of the Suppe fold-bend model was pre- trast to fault-bend folds, fault-propagation folds do not require transfer
sented by Johnson and Berger (1989) (Fig. 9) and Hardy (1995). In the of volume out of the structure (Mitra, 1990).
Johnson and Berger (1989) model the thrust sheet is divided into
three velocity domains, which are related to the flat-ramp-flat geome- 4.4.1. Characteristics
tries of the fault. The velocity domains are separated by velocity discon- The typical characteristics of fault-propagation folding were summa-
tinuities that correspond to the kink-bands A and B in the model of rized by Suppe and Medwedeff (1990) and Shaw et al. (2005). Such
Suppe (1983). In domain 1 the velocity vectors are parallel to the folding is generally located in the hanging wall of a low-angle thrust
lower flat, in domain 2 they are parallel to the ramp and in domain fault. The fold tends to tighten towards the fault, which indicates
3 the velocity vectors are parallel to the upper flat (Fig. 6). This model that fold and fault are related. Often the fold has a distinct asymmetry
perfectly reproduces the structural style of simple fault-bend folds. with a steep or even overturned forelimb and a less steep backlimb
(Fig. 11). The forelimb is narrow, whereas the backlimb is wide. As a
consequence the vergence of the fold points towards the transport
Domain II direction of the thrust. In general, the folds are associated with low-
Domain I Domain III angle thrust faults, which have low displacements (McNaught and
Mitra, 1993). Typically the thrust dies out in the core of the fold. During
VII = V0 propagation of the fault, the fold grows and preserves its initial geome-
VZ
VI = V0 VIII
V0 VX γ try (Shaw et al., 2005). The deformation in the fold is controlled by the
fault angle and the form of the fold (McConnell, 1994). Ongoing defor-
Θ
90 - _
mation can lead to modifications of the fault-propagation fold geometry
Θ 2
90 - _
2 by breakthrough thrusting (Mercier et al., 1997).
Θ

4.4.2. Kinematic models


Fig. 9. Velocity model for fault-bend folds. The model has three velocity domains. The
Kinematic models for fault-propagation folds have been developed
velocity vectors are always parallel to the fault trace. by Suppe and Medwedeff (1990), Mitra (1990) and Mitra and Mount
Modified after Johnson and Berger (1989). (1998). The classic kinematic approach is the parallel kink-fold model
362 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

α'
η
α

α'

Fig. 10. Fault-parallel flow. Material nodes move, parallel to the fault, along flow paths. A material line (l0) is transported by the length z over a concave ramp (angle 2η), which is bisected
by a flow deflector (fd). The line l0 undergoes strain (l1) and the angle α is now α′. The parallel line (l1′) is moved over a similar, but convex ramp, again by the length z. The line is restored
in length to l0 and the angle α′ is restored to α.
Modified after Ziesch et al. (2014).

introduced by Suppe and Medwedeff (1990). This model is similar to models have shown that faults in fold-and-thrust belts can have a
the approach of Suppe (1983) for fault-bend folds. more complex propagation pattern, whereby a fault nucleates within
The basic assumption is a bedding-plane-parallel detachment. The the deforming material and propagates upward and downward, finally
folding is caused when the fault steps up over a ramp. The evolution merging with the basal detachment to form a flat-ramp geometry
of anticlines is controlled by the formation of two kink bands. The (Storti et al., 1997). Tavani et al. (2006) presented a kinematic model
right axial surface of the lower kink band in Fig. 11 (named B in Suppe called double-edge fault-propagation folding that implements these
and Medwedeff, 1990) terminates at the point where the fault steps observations and is able to overcome the limitations of previous models.
up and the left axial surface of the upper kink band (named A′ in The double-edge fault-propagation approach assumes a basal detach-
Suppe and Medwedeff, 1990) ends at the fault tip. Another important ment above which layer-parallel shortening is accommodated (Tavani
plane is the surface AB′, which represents the axial surface of the fold and Storti, 2011). A reverse fault nucleates in the deforming material
(Fig. 11). It starts from the center of the fault trace and terminates at above the basal detachment (Tavani et al., 2006). The key feature
the point where A and B′ merge. The tip of the fault and the branch of of this model is the reverse fault that propagates in two directions,
the axial surface AB′ are in the same stratigraphic position. When the upwards and downwards. Above this reverse fault hanging-wall defor-
fault propagates, the beds roll from a flat position into the steep kink mation occurs. Ongoing shortening leads to continued ramp propaga-
band A–A′ (Suppe and Medwedeff, 1990). Therefore fold growth is a tion and finally the connection of the downward propagating tip with
self-similar process. the basal detachment (Tavani et al., 2006). The double-edge fault-
Jamison (1987) developed a model that extends the parallel fold propagation folding approach is also suitable to model extensional
approach. In this model fault-parallel shear is possible. This is appropri- fault-related folds (Tavani et al., 2006) and the evolution of growth
ate to explain the tightening of fault-propagation folds that is observed strata patterns (Tavani et al., 2007).
in many natural examples. Another kinematic model for fault- Storti and Salvini (1996) proposed another kinematic model
propagation folds was presented by Mitra (1990). This model is based for fault-propagation folding called progressive roll-over fault-
on area-balancing and is suitable to describe the evolution of the fold propagation folding. The aim of this approach is to compensate the
in the case where the fault continues to propagate through undeformed problems of the standard fault-propagation to produce overturned to
units and the initial fault-propagation fold is subsequently translated recumbent forelimbs, which are common in fold-and-thrust belts
into a fault-bend fold as it passes over the footwall ramp (Mitra, (Storti and Salvini, 1996). The model of Storti and Salvini (1996)
1990). Spang and McConnell (1997) showed a kinematic model for extends the approach of Suppe and Medwedeff (1990) by the incorpo-
fault-propagation folds, where deformation occurs in both the hanging rating additional bending of the material in the hanging wall at the
wall and footwall of the faults. This model can also produce a footwall tip line. This requires more axial surfaces than in the model of Suppe
syncline and thus it is a step towards more realistic kinematic models. and Medwedeff (1990) and splits the forelimb into a forward-dipping,
The kink-band migration model of fault-propagation folding pro- a recumbent, and an overturned sector.
duces folds, in which the backlimb dip is parallel to the dip of the
ramp (Fig. 11). However, natural examples often show diverging dips 4.4.3. Velocity models
of ramp and backlimb (Tavani et al., 2006). Furthermore, analogue Waltham (1989) showed that tectonic deformation can be modeled
using a velocity description and subsequently, Waltham and Hardy
(1995) presented velocity solutions for simple shear, pure shear, flex-

A ure, isostasy, and compaction. The advantage of velocity descriptions
B
A´ are their applicability to many different deformation styles and the
γ1 γ1 γ
γ great flexibility of the method to model combined effects of e.g. tecton-
γ γ
ics and sedimentation (Waltham and Hardy, 1995). A velocity field
provides the rate and movement direction of every point in the model
at all times and therefore allows the position of any element in the
model to be tracked. Because the velocity components of all the defor-
mation mechanisms can be added together, this approach is highly
γ1 γ1 Θ AB´ flexible (Waltham and Hardy, 1995). The relevance to this study is
given by the work of Hardy and Poblet (1995), who applied velocity
Fig. 11. Geometry and kinematics of a fault-propagation fold (based on Suppe and descriptions to fault-bend folds and fault-propagation folds, and by
Medwedeff, 1990; Hardy and Poblet, 1995). See text for explanation. the work of Waltham (1989), who used a velocity approach to simulate
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 363

the hanging-wall deformation above listric normal faults. Hardy and and the propagation-to-slip ratio (Allmendinger et al., 2004). The
Connors (2006) extended the velocity description to shear bend folding. propagation-to-slip ratio is one of the most important controlling fac-
tors for the geometry of fault-propagation folds (e.g. McNaught and
4.4.4. Trishear kinematics Mitra, 1993; Hardy and Ford, 1997; Allmendinger, 1998). The value
As seen above, the kinematic model of Suppe and Medwedeff (1990) of the propagation-to-slip ratio is largely responsible for the shape of
produces folds with straight limbs and tight hinges that can be classified hanging-wall anticlines (Allmendinger and Shaw, 2000). The most
as kink or chevron folds. This kink-fold model is limited to reproduce obvious advantage of trishear is the great flexibility in the choice of dif-
several characteristics of fault-propagation folds, such as the curved ferent propagation-to-slip ratios (Allmendinger et al., 2004). Trishear
fold shapes and the presence of footwall synclines, as well as systematic allows the simulation of the strain distribution within fault-related
variations in the thickness and dip of syn-tectonic strata deposited on folds. In cases where strain leads to fracturing, this can help to identify
the anticlinal forelimbs (Allmendinger et al., 2004). This problem was fractures reservoirs and predict strain paths (Allmendinger et al.,
solved by Erslev (1991) by the use of trishear kinematics to explain 2004). With trishear kinematics it is also possible to predict the position
the development of fault-propagation folds. of blind faults from the geometry of the hanging-wall anticline (Bump,
Trishear kinematics are consistent with the thinning of beds in the 2003). The inverse grid search tool of Zehnder and Allmendinger
anticline hinge and thickening of beds adjacent to the syncline hinge. (2000) makes it possible to calculate the best-fit model for the tip line
Hardy and Ford (1997) showed that in contrast to kink-band migration position of the fault, based on the fold shape. In recent years the trishear
models, trishear kinematics produce folds with smooth profiles and approach has been widely applied to natural examples (Hardy and Ford,
rounded hinges. The trishear concept is a numerical approach that as- 1997; Cristallini and Allmendinger, 2001; Cardozo, 2005) and further
sumes a triangular shear zone in front of the tip of a propagating fault computer programs that use trishear kinematics have been developed
(Fig. 12) (Erslev, 1991). At the top of this zone the slip vectors are (Liu et al., 2012). The most recent advances are the realization of
equal to that of the hanging wall, whereas at the base the slip is zero. trishear in 3D (Cristallini et al., 2004; Cardozo, 2008) and the optimiza-
In the trishear zone the slip vectors vary linearly in their magnitude tion of the inverse trishear approach (Cardozo and Aanonsen, 2009;
and orientation from top to bottom (Allmendinger, 1998). In the follow- Cardozo et al., 2011).
ing years, the trishear concept has been extended to a velocity descrip-
tion (Hardy and Ford, 1997; Zehnder and Allmendinger, 2000; Hardy 5. Folding mechanisms, distribution of deformation and
and Allmendinger, 2011). The velocity field is calculated in a Cartesian growth-strata evolution
coordinate system. The ζ–η coordinate systems is fixed to the initial
tip of fault, the X–Y coordinate system is attached to the tip of the There are two major folding mechanisms that occur in the field of
fault and moves with the fault tip (Fig. 12; Zehnder and Allmendinger, fault-related folding: folding by kink-band migration (also known as
2000; Allmendinger et al., 2012). A basic boundary condition is that active-hinge folding), and folding by progressive limb rotation (also
the hanging wall is a rigid block, in which all points move with the known as fix-hinge folding) (Shaw et al., 2005). These two folding
same velocity parallel to the fault (comparable to fault-parallel flow, mechanisms strongly influence the distribution of strain in the fold
see above). The footwall is fixed and the velocity vectors are zero and the geometry of syn-tectonic sediments in the hanging wall of an
(Fig. 12). In addition, area is conserved in the trishear zone (Zehnder evolving anticline. In the case of kink-band migration, material moves
and Allmendinger, 2000; Hardy and Allmendinger, 2011), however across the axial surfaces and the anticlinal limbs widen, but maintain a
bed thickness is not preserved, which means layer-parallel shortening constant dip (Suppe et al., 1992). This can be clearly seen in the case
occurs. This must be considered when line-length balancing is carried of fault-bend folding (Fig. 13). The axial surfaces in the pre-growth stra-
out. ta are parallel. Migration of material across the axial surfaces allows a
The geometry of a trishear fold is determined by the ramp angle, the progressive extension of the rock volume. By the final stage of folding,
slip along the fault, the position of the tip line (X/Y), the trishear angle deformation is distributed throughout the limb of the fault-bend fold

velocity vector

v0 trishear zone
hanging wall x
y (with constant
velocity)

actual
fault tip
1
propagation distance 2
footwall (fixed)

v0 trishear
angle
η ζ
ζ

initial ramp
fault tip angle

Fig. 12. Concept of the trishear approach. Trishear kinematics assumes a triangular shear zone in front of the tip of a propagating fault. At the top of this zone the slip vectors are equal to
that of the hanging wall, whereas at the base the slip is zero. Across the trishear zone, the velocity decreases. Material is deformed within the trishear zone.
Modified after Allmendinger (1998), Zehnder and Allmendinger (2000), Hardy and Allmendinger (2011).
364 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

A) fault-bend fold (limb widening) A) detachment fold (limb rotation)

growth
strata

B) B)

C) narrowing kink-band C) strong deformation in the hinge area


widening backlimb

moderate limb deformation

Fig. 13. Comparison of kink-band migration folding and fixed-hinge folding. In the kink-band migration model (A), the limbs widen; in case of a fixed hinge (B), the limbs rotate. This
affects the related growth-strata. In kink-band migration folding, a growth triangle forms, whereas in the fixed-hinge case, the growth-strata beds also rotate. Based on Shaw et al.
(2005) and Salvini and Storti (2004).

(Salvini and Storti, 2004). Syn-tectonic sedimentation that exceeds the wall strata that are bent down at the fault plane (roll-over anticline),
uplift of the growing anticline is characterized by a narrowing kink- d) transverse folds caused by along-strike changes in fault displace-
band that has converging axial surfaces (Suppe et al. 1992; Shaw et al., ment, e) fault-propagation folds above the tip line of a normal fault
2005). In the case of limb rotation, the hinge is fixed and the limbs and f) differential compaction over a pre-existing fault-scarp. Exten-
rotate, which means the limb dip progressively increases (Fig. 13; sional forced folds can be also added to this list.
Poblet and McClay, 1996; Shaw et al., 2005). The same rock volume The recognition of fault-propagation folding in extensional regimes
stays within the axial surfaces and undergoes progressive deformation. can explain the presence of monoclines and overturned beds (Sharp
In case of fixed-hinge folding, the fold has a narrow, strongly deformed et al., 2000), as well as the occurrence of hanging-wall synclines
area in the axial surface sector, with highest strain at the pin of the (Khalil and McClay, 2002).
axial surface, which decreases towards the limbs (Salvini and Storti, Kinematic models for extensional fault-propagation folds have been
2004). Fixed-hinge folding is an important mechanism in the case also developed based on the trishear approach (Hardy and McClay,
of detachment-folding, as shown by Poblet and McClay (1996). The 1999; Jin and Groshong, 2006). Fig. 14 shows a comparison of contrac-
growth strata evolution reflects the progressive limb rotation; therefore tional and extensional trishear folding. As in the compressional fault-
the limb dip shows a characteristic fanning. The syn-tectonic sediments propagation case (Fig. 14A), a triangular zone above the tip line of the
thin towards the anticline. Due to the limb rotation, the axial surfaces (in this case) normal fault is assumed that is characterized by a velocity
change their dip and some material also moves across the axial surfaces field that decreases from the hanging-wall to the footwall (Jin and
(Shaw et al., 2005). Groshong, 2006). This velocity field determines the strain rate in the
trishear zone. As a consequence of fault-propagation, a monocline
6. Fault-related folding in extensional domains evolves above the tip line of the normal fault (Fig. 14B). The geometry
of the fold and the strain distribution is strongly controlled by the
Fault-related folds have also been recognized in extensional regimes. propagation-to-slip ratio of the fault (Jin and Groshong, 2006). This
They have been analyzed in the field (Howard and John, 1997; Corfield is comparable to the compressional fault-propagation folds. High P/S
and Sharp, 2000; Sharp et al., 2000), with analogue models (Withjack ratios produce narrow folds that are less deformed. The trishear angle
et al., 1990), and numerical simulations (Khalil and McClay, 2002; Jin is an important controlling factor for the geometry of the hanging-wall
and Groshong, 2006). Six different types of extensional fault-related syncline. Large trishear angles cause only small amounts of footwall
folds were summarized by Khalil and McClay (2002): a) hanging-wall deformation and wide hanging-wall synclines, whereas small trishear
fault-bend folds formed due to variations in fault dip, b) “normal” angles produce the opposite (Hardy and McClay, 1999; Khalil and
drag folds developed as a consequence of frictional resistance on a McClay, 2002). The trishear approach allows the simulation of asymmet-
normal fault, with hanging-wall strata dragged up and footwall strata ric extensional fault-propagation folds, which resemble the geometry of
dragged down the fault surface, c) “reverse” drag folds with hanging- field examples (Khalil and McClay, 2002; Jin and Groshong, 2006).
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 365

A) B)

contraction extension

trishear zone trishear zone

Fig. 14. Comparison of trishear in A) compressional and B) extensional regimes. The fault dip is 50°, trishear angle is 60°, and the P/S ratio is 2 in both cases. The slip increment is equal.
Simulation was carried out using the software FaultFoldForward by R. Allmendinger.

7. Fault-related folding and the structure of deforming wedges ramps up to the surface. Variations in décollement dip also control
whether the fault-related folds will be exhumed and eroded (flat
Fault-related folds are dominating elements of fold-and-thrust décollement) or buried under syntectonic sediments (steep hinterland-
belts (e.g. Boyer, 1986; Vann et al., 1986) and accretionary wedges dipping décollement) (Mariotti and Doglioni, 2000). In addition, there
(e.g. Biju-Duval et al., 1982; Gulick et al., 2004). In general, the overall is also interplay of fold uplift, regional subsidence, and sedimentation
geometry of such deforming wedges can be described by the critical- rate (Doglioni and Prosser, 1997). The total fold uplift is the fold uplift
taper wedge theory (Davis et al., 1983; Dahlen, 1984; Dahlen et al., rate minus the regional subsidence. When the uplift rate of the fold is
1984), in which fault-related folds are largely responsible for the inter- higher than the subsidence rate, the fold will be eroded and the onlap
nal structure. of the syn-tectonic strata will move away from the anticlinal crest. If
Trains of fault-bend folds in contractional regimes, such as delta the total fold uplift is negative and the sedimentation rate is high enough,
toe areas and accretionary wedges, are often dominated by ramps the syn-tectonic strata will move towards the anticlinal crest (Doglioni
that sole out into a single décollement layer (Zoetemeijer and Sassi, and Prosser, 1997). A comparable interplay can be observed along
1992; Corredor et al., 2005). The position of the initial décollement is normal faults. In such cases, the total subsidence is the sum of fault-
controlled by pore fluid pressure, the strength of the rocks, and the controlled subsidence and the regional subsidence; it can be positive or
accretion mode of the deforming wedge (Bigi et al., 2003). The position negative. In addition, the subsidence rate can be higher or lower than
of the initial décollement defines the area of the deforming wedge. the sedimentation rate. The interplay of fault slip, regional subsidence,
However, the depth of the basal décollement can vary along strike sedimentation rate, and sea-level fluctuations determine the architec-
(Bigi et al., 2003). The décollement dip has a strong impact on the geom- ture of growth-strata along normal faults (Doglioni et al., 1998). This
etry of fault-related folds in the hanging wall. Mariotti and Doglioni concept is applicable to the sedimentation above roll-over anticlines
(2000) showed that the distance between the ramps increases with in- along normal faults.
creasing dip of the basal décollement. On the other hand, the distance
between the ramps is a controlling factor for the geometry and kinemat- 8. Fault-related folding and inversion structures
ics of piggy-back basins on top of the wedge (Doglioni et al., 1999a). The
décollement dip can be also a function of the orientation of the related Inversion is defined as the reactivation of a normal fault in a reverse
subduction zone. Doglioni et al. (1999b) showed that in westward- sense. Thus inversion structures are characterized by a two or more-
directed subduction zones, the décollement is downwarped, while in phase fault evolution. In general, initial normal movement is followed
east to northeast-directed subduction zones, the related décollement by reverse movement along the same fault. The resulting hanging wall
366 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

is an antiformal feature that is called a “positive inversion structure” the velocity field of kinematic models cannot incorporate effects such
(Williams et al., 1989), “harpoon”, or “arrow head” structure (McClay, as strain partitioning. Finite element modeling and finite difference
1992; McClay and Buchanan, 1992; McClay, 1995). Inversion structures modeling studies that implement, e.g. the mechanical stratigraphy,
have been described by various authors from different locations can close this gap and allow the analysis of the related effects on fault
(Roberts, 1989; Letouzey et al., 1990; Lowell, 1995), and they are char- and fold evolution (e.g. Strayer and Hudleston, 1997; Doglioni et al.,
acteristic of complex intracontinental basins (Kockel, 2003; Mazur et al., 2011; Albertz and Lingrey, 2012; Albertz and Sanz, 2012; Smart et al.,
2005). Inversion structures can occur on a kilometer- (Letouzey et al., 2012). Discrete element modeling is also an appropriate tool to analyze
1990) to meter-scale (Brandes et al., 2012). A comprehensive review the impact of material properties like the mechanical stratigraphy
of inversion structures is given by Turner and Williams (2004). on the deformation style (Hardy and Finch, 2007). Finch et al. (2003)
Inversion structures are further examples of fault-related folding. demonstrated the effects of the cover strength on the geometry of
Their antiformal geometry resembles the shape of fault-propagation fault-propagation folds. In this context a weak cover causes wide
folds with a typically steep forelimb and a gently-dipping backlimb. zones of deformation with limited fault propagation, while a cover
Anticlines can be formed in two ways along an inversion structure. with a high strength results in narrow zones of deformation and pro-
Initial normal faulting may cause the formation of anticlinal roll-overs nounced fault propagation (Finch et al., 2003).
(Yamada and McClay, 2004). The subsequent reverse movements pro- It is immanent that the physical process of generating a kinematic
duce a hanging-wall inversion anticline. Yamada and McClay (2004) model requires fixing the model geometry, i.e. the geometry of fault,
showed that the shape of these anticlines is controlled by the geometry stratigraphic units, and other markers. The modeling itself then will
of the fault plane (concave-up/convex-up). Convex-up listric faults pro- add further constraints to predict the (future or extended) fault
duce less asymmetrical hanging-wall anticlines than the concave-up geometry.
listric faults. Kinematic models of fault-propagation folding, fault-bend folding,
and detachment folding tend to oversimplify the natural deformation
9. Discussion processes (Storti et al., 1997). With scaled sand-box models, Storti
et al. (1997) were able to show that during the formation of a thrust-
Kinematic models of fault-related folding evolved from the early related anticline, the kinematic regime changes from layer-parallel
1980s onward. The ignition was the development of the kink-band shortening, through detachment-folding and fault-propagation folding,
migration models (Suppe, 1983; Suppe and Medwedeff, 1990), which to fault-bend folding. This shows that the mechanism of fault-related
then were followed by numerical descriptions (e.g. Waltham, 1989; folding may change over time (Storti et al., 1997) and this should be
Erslev, 1991; Hardy and Poblet, 1995). All kinematic models of fault- considered in the setup of kinematic models (see below “the future of
related folding should fulfill the general criteria for models in Earth fault-related folding kinematic models”).
Sciences as defined by Greenwood (1989). Models can be validated by Furthermore, all faults and folds have three dimensional termina-
testing their predictions, although it must be kept in mind that models tions and therefore 3D kinematic models are required to cope with
are often not unique and different models may produce the same geom- natural variations in fault slip and fold geometry. At present, only
etry (e.g. Thorbjornsen and Dunne, 1997). the trishear method (Cristallini et al., 2004; Cardozo, 2008) can fulfill
Since their development the advantages of these kinematic models this criterion. Other methods use parallel 2D sections to achieve a
have become apparent. They give insight into deformation processes kinematically-unraveled 3D geometrical model. For instance fault-
on different scales (Yielding et al., 1996; Lohr et al., 2008) and they parallel flow in 3D can only maintain bed length in the transport
allow the prediction of parameters, such as the fault geometry, which direction, but not in any other section that is oblique to the transport
are not usually exposed (Davis and Namson, 1994; Allmendinger and direction (Tanner et al., 2003; Ziesch et al., 2014).
Shaw, 2000). Their simplistic approach can isolate important controlling Fold models are an important tool for predicting strain and thus the
factors e.g. the propagation-to-slip ratio, as demonstrated by Hardy and orientation and distribution of small-scale deformation features such as
Ford (1997) and Allmendinger (1998). Numerical kinematic models are fractures (e.g. Fischer and Wilkerson, 2000; Salvini and Storti, 2004).
also able to describe different deformation styles (Waltham and Hardy, They therefore have a predictive capability (Greenwood, 1989). Never-
1995) and allow a great flexibility in the choice of the input parameters theless, the results strongly depend on which model is used. In case of
(Allmendinger et al., 2004). Using kinematic models of fault-related fixed-hinge folding, the deformation is concentrated in the crest of the
folding can significantly reduce the uncertainties in the reflection seis- anticline, whereas for active hinge folding it is manifested on the
mics workflow by predicting fault trace from fold geometry and tip limbs (Salvini and Storti, 2001). This stresses the demand for unified
line location from the propagation-to-slip ratio. kinematic models of fault-related folding.
Besides these advantages, there are also limitations of the kine-
matic models. First of all, they mostly lack a mechanical foundation
(Allmendinger et al., 2004). The mechanical stratigraphy of a succession 10. The future of fault-related folding kinematic models
has a distinct impact on the deformation style (e.g. Fischer and Jackson,
1999). Hardy (2011) pointed out that velocity models like trishear can Though much work was carried out on fault-related folding during
only describe large-scale structures and the bulk strain, whereas the the last century, there is still potential for further research. Future
distinct element method is able to reproduce small-scale features and work should focus on integrating seismic-scale and outcrop-scale
allow a more detailed strain analysis. In addition, the lack of rheology observations to predict the orientation of features like joint sets or
in the model means that aspects of temporal evolution are not con- deformation bands. This would require the development of coupled
sidered. In case of multi-layer folds, several parameters such as the kinematic and geomechanical models.
material properties of the individual layers, the interlayer cohesion, Understanding small-scale structural elements like joints and defor-
contrasts in the layer properties, and the bulk moduli of the whole mations bands is crucial to develop more comprehensive models that
layer package control the behavior during deformation (Woodward, consider the influence of structural elements on the sub-surface fluid
1997). Johnson and Johnson (2002) have shown that the shape of the flow. Increasing energy demands makes it imperative to achieve better
forelimb of forced folds is strongly controlled by the rheology of the reservoir characterization in hydrocarbon geology and geothermal
cover but only weakly controlled by the geometry of the fault. applications. Kinematic fault-related folding models are a strong tool
With true geomechanical models it should be possible, for instance, to predict the orientation of small-scale tectonic fabrics, as demonstrat-
to simulate parallel and similar fold morphologies, or predict the path- ed by Lohr et al. (2008) and Brandenburg et al. (2012). Further research
ways of new faults. Due to the need to maintain a physical continuum, on fault-related folding should also focus on the evolution of soft
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 367

A) fault-parallel flow
fixed pin
loose pin

B) trishear
A
sheared fixed pin
loose pin

Fig. 15. Fault-related folds produced by A) fault-parallel flow over a detachment and ramp of 30°, B) trishear of a ramp propagating upwards from a detachment at 30°, P/S ratio 2. The slip
increment is equal. The resulting hanging-wall folds are substantially different in their geometry, which is a function of the P/S ratio. Future algorithms could aim to combine these
approaches to create a unified kinematic model.

sediment fault-related folds to generate holistic models for mass are the presence of layer-parallel strain and mechanical stratigraphy.
wasting along continental margins. Layer-parallel strain is important for the evolution of detachment
Fault geometry is more complicated than has been modeled up folds and fault-bend faults, where fixed stratigraphic detachments are
to now. Using natural fault geometries would boost the reliability of involved. The mechanical stratigraphy determines the position of de-
the results and lead to the development of new algorithms. Steps into tachments and the development of ramps, which are the key elements
this direction were recently made by Brandenburg (2013) who extend- for fault-bend folds and fault-propagation folds. Mechanical stratigra-
ed the trishear algorithm to naturally-curved fault planes. phy is also significant for the development of detachment folds because
All kinematic models of fault-related folding in the present-day they require a stack of less competent and competent rock units.
include similar simplifications and limitations, such as: Kinematic models clearly enhanced the understanding of the defor-
mation processes that control fault-related folds. However, this does not
1. They enforce the same kinematic behavior throughout the model. If
mean that fold geometry is uniquely related to fault geometry; different
they could be combined, for instance using 40% trishear and 60%
kinematic approaches can lead to the same fold shape.
fault-parallel flow, this would allow the chance to get closer to natu-
ral deformation (Fig. 15).
Acknowledgments
2. They are mostly two dimensional. Natural deformation is always
three-dimensional. Two-dimensional models can only model plane
This manuscript could not be possible without the discussion and
strain, which is exceptional in nature. All present-day algorithms
suggestions of many colleagues over many years. We would like to
should attempt to make the huge leap to three dimensions. The
thank P. Blisniuk, B. Leiss, K. Reicherter, A. Vollbrecht, and J. Winsemann
result would be far more realistic, albeit at the cost of computing
for discussions. We are grateful to F. Storti, E. Tavarnelli and an anony-
power.
mous reviewer for their constructive comments and C. Doglioni for
There are a number of general ways that the algorithms could be helpful comments and editorial work. Midland Valley Exploration Ltd.
improved or even new algorithms written. For instance: is gratefully acknowledged for the use of their software package Move.

1. Analogue models, using, for instance, plasticine or sand, could be


References
used, maybe analyzed using, e.g. particle image velocimetry, to
design new velocity models for fault movement. The results could Aikulola, U.O., Olotu, S.O., Yamusa, I., 2010. Investigating fault shadows in the Niger Delta.
then be converted to numerical algorithms for kinematic modeling. Lead. Edge 29, 16–22.
Albertz, M., Lingrey, S., 2012. Critical state finite element models of contractional fault-
2. Ground-truthing. The authors are of the opinion that far more related folding: part 1. Mechanical analysis. Tectonophysics 576–577, 133–149.
emphasis should be placed on the analysis of natural structures Albertz, M., Sanz, P.F., 2012. Critical state finite element models of contractional fault-
(e.g. Tanner et al., 2010). Algorithms can then be discriminated as related folding: part 2. Mechanical analysis. Tectonophysics 576–577, 150–170.
Allmendinger, R.W., 1998. Inverse and forward numerical modeling of trishear fault-
to their applicability to particular structures. An algorithm, however propagation folds. Tectonics 17, 640–656.
good, is, without ground-truthing, of much less value. Allmendinger, R.W., Shaw, J.H., 2000. Estimation of fault propagation distance from fold
3. Extraction of material properties from geophysical methods, such as shape: implications for earthquake hazard assessment. Geology 28, 1099–1102.
Allmendinger, R.W., Zapata, T., Manceda, R., Dzelalija, F., 2004. Trishear kinematic model-
reflection seismics. Examples of this include Poisson's Ratio from the
ing of structures, with examples from the Neuquén Basin, Argentina. In: McClay, K.R.
seismic reflection velocities, i.e. Vp/Vs (e.g. Gassaway and Richgels, (Ed.), Thrust Tectonics and Hydrocarbon Systems AAPG Mem. 82, 356–371.
1983; Beilecke et al., 2014). This will allow rheological parameters Allmendinger, R.W., Cardozo, N., Fisher, D.M., 2012. Structural Geology Algorithms —
Vectors and Tensors. Cambridge University Press (289 pp.).
to be entered in to models.
Alsop, G.I., Marco, S., 2012. A large-scale radial pattern of seismogenic slumping towards
the Dead Sea Basin. J. Geol. Soc. Lond. 169, 99–110.
11. Conclusions Apotria, T., Wilkerson, M.S., 2002. Seismic expression and kinematics of a fault-related
fold termination: Rosario structure, Maracaibo Basin, Venezuela. J. Struct. Geol. 24,
671–687.
Fault-related folding is one of the most important deformation Aydin, A., 2000. Fractures, faults, and hydrocarbon entrapment, migration and flow. Mar.
processes in structural geology at all scales and occurs in very different Pet. Geol. 17, 797–814.
tectonic settings (compressional, extensional and strike-slip). Never- Becker, A., 1995. Conical drag folds as kinematic indicators for strike-slip fault motion.
J. Struct. Geol. 17, 1497–1506.
theless, the evolution of the structures can be explained or be at least Beilecke, T., Ziesch, J., Tanner, D.C., Krawczyk, C.M., the Protect Research Group, 2014.
partly reconstructed by simple kinematic models. Basic prerequisites Poisson's ratio model derived from P- and S-wave reflection seismic data at
368 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

the CO2CRC Otway Project pilot site, Australia. European Geosciences Union General Culshaw, M.G., 2005. From concept towards reality: developing the attributed 3D geological
Assembly 2014, Vienna, Austria, 27 April–2 May 2014. model of the shallow subsurface. Q. J. Eng. Geol. Hydrogeol. 38, 231–284.
Berger, P., Johnson, A.M., 1980. First order analysis of deformation of a thrust sheet Daëron, M., Klinger, Y., Tapponnier, P., Elias, A., Jacques, E., Sursock, A., 2007. 12,000-
moving over a ramp. Tectonophysics 70, 9–24. year-long record of 10 to 13 paleoearthquakes on the Yammoûneh fault, levant
Bernard, S., Avouac, J.-P., Dominguez, S., Simoes, M., 2007. Kinematics of fault-related fault system, Lebanon. Bull. Seismol. Soc. Am. 97, 749–771.
folding derived from sandbox experiments. J. Geophys. Res. 112, B03S12. Dahlen, F.A., 1984. Noncohesive critical Coulomb wedges: an exact solution. J. Geophys.
Bigi, S., Lenci, F., Doglioni, C., Moore, J.C., Carminati, E., Scrocca, D., 2003. Décollement Res. 89, 10125–10133.
depth vs accretionary prism dimension in the Apennines and the Barbados. Tectonics Dahlen, F.A., Suppe, J., Davis, D., 1984. Mechanics of fold-and-thrust belts and accretionary
22, 1010. wedges: cohesive Coulomb theory. J. Geophys. Res. 89, 10087–10101.
Biju-Duval, B., Le Quellec, P., Mascle, A., Renard, V., Valery, P., 1982. Multibeam bathymet- Dahlstrom, C.D.A., 1969. Balanced cross sections. Can. J. Earth Sci. 6, 743–757.
ric survey and high resolution seismic investigations on the Barbados Ridge Complex Dahlstrom, C.D.A., 1990. Geometric constraints derived from the law of conservation of
(eastern Caribbean): a key to the knowledge and interpretation of an accretionary volume and applied to evolutionary models for detachment folding. AAPG Bull. 74,
wedge. Tectonophysics 86, 275–304. 336–344.
Biot, M.A., 1961. Theory of folding of stratified visco-elastic media and its implications in Davis, T.L., Namson, J.S., 1994. A balanced cross-section of the 1994 Northridge earth-
tectonics and orogenesis. GSA Bull. 72, 1595–1620. quake, southern California. Nature 372, 167–169.
Boyer, S.E., 1986. Styles of folding within thrust sheets: examples from the Appalachian Davis, D., Suppe, J., Dahlen, F.A., 1983. Mechanics of fold-and-thrust belts and accretionary
and Rocky Mountains of the USA and Canada. J. Struct. Geol. 8, 325–339. wedges. J. Geophys. Res. 88, 1153–1172.
Brandenburg, J.P., 2013. Trishear for curved faults. J. Struct. Geol. 53, 80–94. Delcaillau, B., Deffontaines, B., Floissac, L., Angelier, J., Deramond, J., Souquet, P., Chu, H.T.,
Brandenburg, J.P., Alpak, F.O., Solum, J.G., Naruk, S.J., 2012. A kinematic trishear model to Lee, J.F., 1998. Morphotectonic evidence from lateral propagation of an active frontal
predict deformation bands in a fault-propagation fold, East Kaibab monocline, Utah. fold; Pakuashan anticline, foothills of Taiwan. Geomorphology 24, 263–290.
AAPG Bull. 96, 109–132. Delcaillau, B., Carozza, J.M., Laville, E., 2006. Recent fold growth and drainage development:
Brandes, C., Le Heron, D.P., 2010. The glaciotectonic deformation of Quaternary sediments the Janauri and Chandigarh anticlines in the Siwalik foothills, northwest India.
by fault-propagation folding. Proc. Geol. Assoc. 121, 270–280. Geomorphology 76, 241–256.
Brandes, C., Astorga, A., Back, S., Littke, R., Winsemann, J., 2007a. Fault controls on DePaor, D.G., 1988. Balanced section in thrust belts part 1: construction. AAPG Bull. 72,
sediment distribution patterns, Limón Basin, Costa Rica. J. Pet. Geol. 30, 25–40. 73–90.
Brandes, C., Astorga, A., Blisniuk, P., Littke, R., Winsemann, J., 2007b. Anatomy of anti- DeVecchio, D.E., Keller, E.A., Fuchs, M., Owen, L.A., 2012. Late Pleistocene structural
clines, piggy-back basins and growth strata: a case study from the Limón fold-and- evolution of the Camarillo fold belt: implications for lateral fault growth and seismic
thrust belt, Costa Rica. In: Nichols, G., Williams, E., Paola, C. (Eds.), Sedimentary hazard in southern California. Lithosphere 4, 91–109.
Processes, Environments and Basins. IAS Special Publication, 38, pp. 91–110. Doglioni, C., Carminati, E., 2008. Structural styles and Dolomites field trip. Mem. Descr.
Brandes, C., Winsemann, J., Roskosch, J., Meinsen, J., Tanner, D.C., Frechen, M., Steffen, H., Carta Geol. Ital. 82, 1–299.
Wu, P., 2012. Activity of the Osning thrust during the late Weichselian: ice-sheet and Doglioni, C., Prosser, G., 1997. Fold uplift versus regional subsidence and sedimentation
lithosphere interactions. Quat. Sci. Rev. 38, 49–62. rate. Mar. Pet. Geol. 14, 179–190.
Braun, J., Sambridge, M., 1994. Dynamical Lagrangian Remeshing (DLR): a new algorithm Doglioni, C., D'Agostino, N., Mariotti, G., 1998. Normal faulting versus regional subsidence
for solving large strain deformation problems and its application to fault-propagation and sedimentation rate. Mar. Pet. Geol. 15, 737–750.
folding. Earth Planet. Sci. Lett. 124, 211–220. Doglioni, C., Merlini, S., Cantarella, G., 1999a. Foredeep geometries at the front of the
Bump, A.P., 2003. Reactivation, trishear modelling, and folded basement in Laramide Apennines in the Ionian Sea (central Mediterranean). Earth Planet. Sci. Lett. 168,
uplifts: implications for the origins of intra-continental faults. GSA Today 2003, 243–254.
4–10 (March). Doglioni, C., Harabaglia, P., Merlini, S., Mongelli, F., Peccerillo, A., Piromallo, C., 1999b.
Burbank, D.W., McLean, J.K., Bullen, M., Abdrakhmatov, K.Y., Miller, M.M., 1999. Orogens and slabs vs. their direction of subduction. Earth-Sci. Rev. 45, 167–208.
Partitioning of intermontane basins by thrust-related folding, Tien Shan, Kyrgyzstan. Doglioni, C., Barba, S., Carminati, E., Riguzzi, F., 2011. Role of the brittle–ductile transition
Basin Res. 11, 75–92. on fault activation. Phys. Earth Planet. Inter. 184, 160–171.
Burke, H., Phillips, E., Lee, J.R., Wilkinson, I.P., 2009. Imbricate thrust stack model for the Dolan, J.F., Christofferson, S.A., Shaw, J.H., 2003. Recognition of paleoearthquakes on the
formation of glaciotectonic rafts: an example from the Middle Pleistocene of north Puente Hills blind thrust fault, California. Science 300, 115–118.
Norfolk, UK. Boreas 38, 620–637. Egan, S.S., Buddin, T.S., Kane, S.J., Williams, G.D., 1997. Three-dimensional modelling and
Buxtorf, A., 1916. Prognosen und Befunde beim Hauensteinbasis- und Grenchenbergtunnel visualisation in structural geology: new techniques for the restoration and balancing
und die Bedeutung der letztern für die Geologie des Juragebirges. Abh. Naturforsch. Ges. of volumes. Proceedings of the 1996 geoscience information group conference on
Basel 27, 181–205. geological visualisation. Electron. Geol. 1, 67–82.
Cadell, H.M., 1889. Experimental researches in mountain building. Trans. R. Soc. Edinb. 35, Elliott, D., 1983. The construction of balanced cross sections. J. Struct. Geol. 5, 101.
337–357. Epard, J.-L., Groshong, R.H., 1993. Excess area and depth to detachment. AAPG Bull. 77,
Calamita, F., Pace, P., Satolli, S., 2012. Coexistence of fault-propagation and fault-bend 1291–1302.
folding in curve-shaped foreland fold-and-thrust belts: examples from the Northern Epard, J.-L., Groshong, R.H., 1995. Kinematic model of detachment folding including limb
Apennines (Italy). Terra Nova 24, 396–406. rotation, fixed hinges and layer-parallel strain. Tectonophysics 247, 85–103.
Camerlo, R.H., Benson, E.F., 2006. Geometric and seismic interpretation of the Perdido fold Erickson, S.G., 1995. Mechanics of triangle zones and passive roof duplexes: Implications
belt: northwestern deep-water Gulf of Mexico. AAPG Bull. 90, 363–386. of finite element models. Tectonophysics 245, 1–11.
Cardozo, N., 2005. Trishear modelling of fold bedding data along a topographic profile. Erickson, S.G., Jamison, W.R., 1995. Viscous–plastic finite-element models of fault-bend
J. Struct. Geol. 27, 495–502. folds. J. Struct. Geol. 17, 561–573.
Cardozo, N., 2008. Trishear in 3D. Algorithms, implementation, and limitations. J. Struct. Erslev, E.A., 1991. Trishear fault-propagation folding. Geology 19, 617–620.
Geol. 30, 327–340. Erslev, E.A., Mayborn, K.R., 1997. Multiple geometries and modes of fault-propagation
Cardozo, N., Aanonsen, S., 2009. Optimized trishear inverse modeling. J. Struct. Geol. 31, folding in the Canadian thrust belt. J. Struct. Geol. 19, 321–335.
546–560. Ferrill, D.A., Morris, A.P., McGinnis, R.N., 2012. Extensional fault-propagation folding in
Cardozo, N., Jackson, C.A.-L., Whipp, P.S., 2011. Determining the uniqueness of best-fit mechanically layered rocks: the case against the frictional drag mechanism.
trishear models. J. Struct. Geol. 33, 1063–1078. Tectonophysics 576–577, 78–85.
Chamberlin, R.T., 1910. The Appalachian folds of central Pennsylvania. J. Geol. 18, Finch, E., Hardy, S., Gawthorpe, R., 2003. Discrete element modelling of contractional
228–251. fault-propagation folding above rigid basement fault blocks. J. Struct. Geol. 25,
Chapman, T.J., Williams, G.D., 1984. Displacement–distance methods in the analysis of 515–528.
fold-thrust structures and linked-fault systems. J. Geol. Soc. 141, 121–128. Finch, E., Hardy, S., Gawthorpe, R., 2004. Discrete element modelling of extensional fault-
Chen, W.-S., Huang, B.-S., Chen, Y.-G., Lee, Y.-H., Yang, C.-N., Lo, C.-H., Chang, H.-C., Sung, propagation folding above rigid basement fault blocks. Basin Res. 16, 489–506.
Q.-C., Huang, N.-W., Lin, C.-C., Sung, S.-H., Lee, K.-J., 2001. 1999 Chi-Chi earthquake: a Fischer, M.P., Jackson, P.B., 1999. Stratigraphic controls on deformation patterns in fault-
case study on the role of thrust-ramp structures for generating earthquakes. Bull. related folds: a detachment fold example from the Sierra Madre Oriental, northeast
Seismol. Soc. Am. 91, 986–994. Mexico. J. Struct. Geol. 21, 613–633.
Cloos, H., 1936. Einführung in die Geologie. Verlag von Gebrüder Bornträger, Berlin (503 Fischer, M.P., Wilkerson, M.S., 2000. Predicting the orientation of joints from fold shape:
pp.). results of pseudo-three-dimensional modeling and curvature analysis. Geology 28,
Contreras, J., 2010. A model for low amplitude detachment folding and syntectonic 15–18.
stratigraphy based on the conservation of mass equation. J. Struct. Geol. 32, 566–579. Ford, M., Williams, E.A., Artoni, A., Vergés, J., Hardy, S., 1997. Progressive evolution of a
Cooper, M.A., Trayner, P.M., 1986. Thrust-surface geometry: implications for thrust-belt fault-related fold pair from growth strata geometries, Sant Llorenc de Morunys, SE
evolution and section-balancing techniques. J. Struct. Geol. 8, 305–312. Pyrenees. J. Struct. Geol. 19, 413–441.
Corfield, S., Sharp, I.R., 2000. Structural style and stratigraphic architecture of fault prop- Gassaway, G.S., Richgels, H.J., 1983. SAMPLE: seismic amplitude measurements for primary
agation folding in extensional settings: a seismic example from the Smorbukk area, lithology estimation. 53rd An. Inter. Meeting Soc. Expl. Geophysics September, Las
Halten Terrace, Mid-Norway. Basin Res. 12, 329–341. Vegas, pp. 610–613.
Corredor, F., Shaw, J.H., Bilotti, F., 2005. Structural styles in the deep-water fold and thrust Grasemann, B., Martel, S., Passchier, C., 2005. Reverse and normal drag along a fault.
belts of the Niger Delta. AAPG Bull. 89, 753–780. J. Struct. Geol. 27, 999–1010.
Cristallini, E.O., Allmendinger, R.W., 2001. Pseudo 3-D modeling of trishear fault- Greenwood, H.J., 1989. On models and modeling. Can. Mineral. 27, 1–14.
propagation folding. J. Struct. Geol. 23, 1883–1899. Groshong, R.H., Epard, J.-L., 1994. The role of strain in area-constant detachment folding.
Cristallini, E.O., Giambiagi, L., Allmendinger, R.W., 2004. True three-dimensional trishear: J. Struct. Geol. 16, 613–618.
a kinematic model for strike-slip and oblique-slip deformation. Geol. Soc. Am. Bull. Gulick, S.P.S., Bangs, N.L.B., Shipley, T.H., Nakamura, Y., Moore, G., Kuramoto, S., 2004.
116, 938–952. Three-dimensional architecture of the Nankai accretionary prism's imbricate thrust
C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370 369

zone off Cape Muroto, Japan: Prism reconstruction via en echelon thrust propagation. Lisle, R.J., 1994. Detection of zones of abnormal strains in structures using Gaussian curva-
J. Geophys. Res. 109, B02105. ture analysis. AAPG Bull. 78, 1811–1819.
Guzofski, C.A., Shaw, J.H., Lin, G., Shearer, P.M., 2007. Seismically active wedge structure be- Liu, C., Yin, H., Zhu, L., 2012. TrishearCreator: a tool for the kinematic simulation and strain
neath the Coalinga anticline, San Joaquin basin, California. J. Geophys. Res. 112, B03S05. analysis of trishear fault-propagation folding with growth strata. Comput. Geosci. 49,
Hamblin, W.K., 1965. Origin of “reverse drag” on the downthrown side of normal faults. 200–206.
GSA Bull. 76, 1145–1164. Lohr, T., Krawczyk, C.M., Oncken, O., Tanner, D.C., Samiee, R., Endres, H., Thierer, P.,
Hansen, E., 1971. Strain Facies. Springer, New York (207 pp.). Trappe, H., Bachmann, R., Kukla, P.A., 2008. Prediction of subseismic faults and
Hardy, S., 1995. A method for quantifying the kinematics of fault-bend folding. J. Struct. fractures: Integration of three-dimensional seismic data, three-dimensional retro-
Geol. 17, 1785–1788. deformation, and well data on an example of deformation around an inverted fault.
Hardy, S., 2011. Cover deformation above steep, basement normal faults: insights from 2D AAPG Bull. 92 (4), 473–485.
discrete element modeling. Mar. Pet. Geol. 28, 966–972. Lowell, J.D., 1995. Mechanics of basin inversion from worldwide examples. In: Buchanan,
Hardy, S., Allmendinger, R.W., 2011. Trishear: a review of kinematics, mechanics, and J.G., Buchanan, P.G. (Eds.), Basin Inversion Geol. Soc. Spec. Publ. 88, 39–57 (London).
applications. In: McClay, K., Shaw, J.H., Suppe, J. (Eds.), Thrust-fault Related Folding Mariotti, G., Doglioni, C., 2000. The dip of the foreland monocline in the Alps and
AAPG Mem. 94, 95–119. Apennines. Earth Planet. Sci. Lett. 181, 191–202.
Hardy, S., Connors, C.D., 2006. Short note: a velocity description of shear fault-bend Marrett, R., Bentham, P.A., 1997. Geometric analysis of hybrid fault-propagation/
folding. J. Struct. Geol. 28, 536–543. detachment folds. J. Struct. Geol. 19, 243–248.
Hardy, S., Finch, E., 2005. Discrete-element modelling of detachment folding. Basin Res. Mazur, S., Scheck-Wenderoth, M., Krzywiec, P., 2005. Different modes of the Late
17, 507–520. Cretaceous/Early Tertiary inversion in the North German and Polish basins. Int.
Hardy, S., Finch, E., 2007. Mechanical stratigraphy and the transition from trishear to J. Earth Sci. 94, 782–798.
kink-band fault-propagation folds forms above blind basement thrust faults. A McClay, K.R., 1992. Glossary of thrust tectonics terms. In: McClay, K.R. (Ed.), Thrust
discrete-element study. Mar. Pet. Geol. 24, 75–90. Tectonics. Chapman and Hall, London, pp. 419–433.
Hardy, S., Ford, M., 1997. Numerical modelling of trishear fault propagation folding. McClay, K.R., 1995. The geometries and kinematics of inverted fault systems: a review
Tectonics 16, 841–854. of analogue model studies. In: Buchanan, J.G., Buchanan, P.G. (Eds.), Basin
Hardy, S., McClay, K., 1999. Kinematic modelling of extensional fault-propagation folding. InversionGeol. Soc. Spec. Publ. 88, 97–118 (London).
J. Struct. Geol. 21, 695–702. McClay, K., 2011. Introduction to thrust fault-related folding. In: McClay, K., Shaw, J.H.,
Hardy, S., Poblet, J., 1994. Geometric and numerical model of progressive limb rotation in Suppe, J. (Eds.), Thrust-fault Related Folding AAPG Mem. 94, 1–19.
detachment folds. Geology 22, 371–374. McClay, K.R., Buchanan, P.G., 1992. Thrust faults in inverted extensional basins. In:
Hardy, S., Poblet, J., 1995. The velocity description of deformation. Paper 2: sediment McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall, London, pp. 93–121.
geometries associated with fault-bend and fault-propagation folds. Mar. Pet. Geol. McConnell, D.A., 1994. Fixed-hinge, basement-involved fault-propagation folds,
12, 165–176. Wyoming. GSA Bull. 106, 1583–1593.
Hardy, S., Poblet, J., McClay, K., Waltham, D., 1996. Mathematical modelling of growth McConnell, D.A., Kattenhorn, S.A., Brenner, L.M., 1997. Distribution of fault slip in outcrop-
strata associated with fault-related fold structures. Modern Developments in scale fault-related folds, Appalachian Mountains. J. Struct. Geol. 19, 57–267.
Structural Interpretation, Validation and Modeling, pp. 265–282. McNaught, M.A., Mitra, G., 1993. A kinematic model for the origin of footwall synclines.
Homza, T.X., Wallace, W.K., 1995. Geometric and kinematic models for detachment folds J. Struct. Geol. 15, 805–808.
with fixed and variable detachment depths. J. Struct. Geol. 17, 575–588. Medwedeff, D.A., 1989. Growth fault-bend folding at the Southeast Lost Hills, San Joaquin
Homza, T.X., Wallace, W.K., 1997. Detachment folds with fixed hinges and variable Valley, California. AAPG Bull. 73, 54–67.
detachment depth, northeastern Brooks Range, Alaska. J. Struct. Geol. 19, 337–354. Medwedeff, D.A., Suppe, J., 1997. Multibend fault-bend folding. J. Struct. Geol. 19,
Hopkins, O.B., 1917. Oil and Gas Possibilities of the Hatchetigbee Anticline, Alabama. 279–292.
Hossack, J.R., 1979. The use of balanced cross-sections in the calculation of orogenic Mercier, E., Outtani, F., Frizon de LaMotte, D., 1997. Late-stage evolution of fault-
contraction: a review. J. Geol. Soc. 136, 705–711. propagation folds: principles and examples. J. Struct. Geol. 19, 185–193.
Howard, K.A., John, B.E., 1997. Fault-related folding during extension: plunging Mitra, S., 1990. Fault-propagation folds: geometry, kinematic evolution, and hydrocarbon
basement-cored folds in the Basin and Range. Geology 25, 223–226. traps. AAPG Bull. 74, 921–945.
Ingram, G.M., Chisholm, T.J., Grant, C.J., Hedlund, C.A., Stuart-Smith, P., Teasdale, J., 2004. Mitra, S., 1993. Geometry and kinematic evolution of inversion structures. AAPG Bull. 77,
Deepwater North West Borneo: hydrocarbon accumulation in an active fold and 1159–1191.
thrust belt. Mar. Pet. Geol. 21, 879–887. Mitra, S., 2002a. Fold-accommodation faults. AAPG Bull. 86, 671–693.
Jackson, J., Norris, R., Youngson, J., 1996. The structural evolution of active fault and fold Mitra, S., 2002b. Structural models of faulted detachment folds. AAPG Bull. 86, 1673–1694.
systems in central Otago, New Zealand: evidence revealed by drainage patterns. Mitra, S., 2003. A unified kinematic model for the evolution of detachment folds. J. Struct.
J. Struct. Geol. 18, 217–234. Geol. 25, 1659–1673.
Jamison, W., 1987. Geometric analysis of fold development in overthrust terranes. Mitra, S., Mount, V.S., 1998. Foreland basement involved structures. AAPG Bull. 82, 70–109.
J. Struct. Geol. 9, 207–219. Mitra, S., Namson, J., 1989. Equal-area balancing. Am. J. Sci. 289, 563–599.
Jamison, W.R., 1996. Mechanical models of triangle zone evolution. Bull. Can. Petrol. Geol. Morley, C.K., King, R., Hillis, R., Tingay, M., Backe, G., 2011. Deepwater fold and thrust belt
44, 180–194. classification, tectonics, structure and hydrocarbon prospectivity: a review. Earth Sci.
Jin, G., Groshong, R.H., 2006. Trishear kinematic modeling of extensional fault- Rev. 104, 41–91.
propagation folding. J. Struct. Geol. 28, 170–183. Mueller, K., Suppe, J., 1997. Growth of Wheeler Ridge anticline, California: geomorphic
Johnson, A.M., Berger, P., 1989. Kinematics of fault-bend folding. Eng. Geol. 27, 181–200. evidence for fault-bend folding behaviour during earthquakes. J. Struct. Geol. 19,
Johnson, K.M., Johnson, A.M., 2002. Mechanical analysis of the geometry of forced-folds. 383–396.
J. Struct. Geol. 24, 401–410. Namson, J.S., Davis, T.L., 1988. Seismically active fold and thrust belt in the San Joaquin
Johnson, K.M., Segall, P., 2004. Imaging the ramp-décollement geometry of the Chelungpu Valley, central California. GSA Bull. 100, 257–273.
fault using coseismic GPS displacements from the 1999 Chi-Chi, Taiwan earthquake. Nelson, R.A., 2001. Geologic Analysis of Naturally Fractured Reservoirs, 2nd edn. Gulf
Tectonophysics 378, 123–139. Publishing Co. Book Division (332 pp.).
Jones, P.B., 1996. Triangle zone geometry, terminology and kinematics. Bull. Can. Petrol. Okamura, Y., Ishiyama, T., Yanagisawa, Y., 2007. Fault-related folds above the source fault
Geol. 44, 139–152. of the 2004 mid-Niigata Prefecture earthquake, in a fold-and-thrust belt caused by
Kao, H., 2000. The Chi-Chi earthquake sequence: active out-of-sequence thrust faulting in basin inversion along the eastern margin of the Japan Sea. J. Geophys. Res. 112,
Taiwan. Science 288, 2346–2349. B03S08.
Keller, E.A., Zepeda, R.I., Rockwell, T.K., Ku, T.L., Dinklage, W.S., 1998. Active tectonics at Okubo, C.H., Schultz, R.A., 2004. Mechanical stratigraphy in the western equatorial region
Wheeler Ridge, southern San Joaquin Valley, California. GSA Bull. 110, 298–310. of Mars based on thrust fault-related fold topography and implications for near-
Kent, W.N., Dasgupta, U., 2004. Structural evolution in response to fold and thrust belt surface volatile reservoirs. GSA Bull. 116, 594–605.
tectonics in northern Assam. A key to hydrocarbon exploration in the Jaipur anticline Peacock, D.C.P., Knipe, R.J., Sanderson, D.J., 2000. Glossary of normal faults. J. Struct. Geol.
area. Mar. Pet. Geol. 21, 785–803. 22, 291–305.
Khalil, S.M., McClay, K.R., 2002. Extensional fault-related folding, northwestern Red Sea, Poblet, J., Hardy, S., 1995. Reverse modelling of detachment folds; application to the
Egypt. J. Struct. Geol. 24, 743–762. Pico del Aguila anticline in the South Central Pyrenees (Spain). J. Struct. Geol. 17,
Kockel, F., 2003. Inversion structures in Central Europe - Expressions and reasons, an open 1707–1724.
discussion. Neth. J. Geosci./Geol. Mijnb. 82, 367–382. Poblet, J., McClay, K., 1996. Geometry and kinematics of single-layer detachment folds.
Korsch, R.J., 2004. A Permian–Triassic retroforeland thrust system — the New England AAPG Bull. 80, 1085–1109.
orogen and adjacent sedimentary basins, eastern Australia. In: McClay, K.R. (Ed.), Poblet, J., McClay, K., Storti, F., Muñoz, J.A., 1997. Geometries of syntectonic sediments
Thrust Tectonics and Hydrocarbon Systems AAPG Mem. 82, 515–537. associated with single-layer detachment folds. J. Struct. Geol. 19, 369–381.
Laubscher, H.P., 1985. Large-scale, thin-skinned thrusting in the southern Alps: Kinematic Ramberg, H., 1963. Evolution of drag folds. Geol. Mag. 100, 97–110.
models. GSA Bull. 96, 710–718. Reches, Z., 1976. Analysis of joints in two monoclines in Israel. GSA Bull. 87, 1654–1662.
Letouzey, J., Werner, P., Marty, A., 1990. Fault reactivation and structural inversion. Reches, Z., Eidelman, A., 1995. Drag along faults. Tectonophysics 247, 145–156.
Backarc and intraplate compressive deformations. Example of the eastern Sunda Resor, P.G., Pollard, D.D., 2012. Reverse drag revisited: why footwall deformation may be
shelf (Indonesia). Tectonophysics 183, 341–362. the key to inferring listric fault geometry. J. Struct. Geol. 41, 98–109.
Lettis, W.R., Wells, D.L., Baldwin, J.N., 1997. Empirical observations regarding reverse Rettger, R.E., 1935. Experiments on soft-rock deformation. AAPG Bull. 19, 271–292.
earthquakes, blind thrust faults, and Quaternary deformation: are blind thrust faults Rich, J.L., 1921. Moving underground water as a primary cause of the migration and
truly blind? Bull. Seismol. Soc. Am. 87, 1171–1198. accumulation of oil and gas. Econ. Geol. 16, 347–371.
Lin, M.L., Wang, C.P., Chen, W.S., Yang, C.N., Jeng, F.S., 2007. Inference of trishear-faulting Rich, J.L., 1934. Mechanics of low-angle overthrust faulting as illustrated by Cumberland
processes from deformed pregrowth and growth strata. J. Struct. Geol. 29, 1267–1280. thrust block, Virginia, Kentucky, and Tennessee. AAPG Bull. 18, 1584–1596.
370 C. Brandes, D.C. Tanner / Earth-Science Reviews 138 (2014) 352–370

Roberts, D.G., 1989. Basin inversion in and around the British Isles. In: Cooper, M.A., Tanner, D.C., Bense, F.A., Ertl, G., 2011. Kinematic retro-modelling of a cross-section
Williams, G.D. (Eds.), Inversion Tectonics Geol. Soc. Spec. Publ. 44, 131–150 (London). through the Western Irish Namurian Basin. In: Poblet, J., Lisle, R.J. (Eds.), Kinematic
Roering, J.J., Cooke, M.L., Pollard, D.D., 1997. Why blind thrust faults do not propagate to evolution and structural styles of fold- and-thrust belts Geol. Soc. Lond. Spec. Publ.
the Earth's surface: numerical modeling of coseismic deformation associated with 349, 61–76.
thrust-related anticlines. J. Geophys. Res. 102 (B6), 11901–11912. Tavani, S., Storti, F., 2006. Fault-bend folding as end-member solution of (double-edge)
Rowan, M.G., 1997. Three-dimensional geometry and evolution of a segmented detach- fault-propagation folding. Terra Nova 18, 270–275.
ment fold, Mississippi Fan foldbelt, Gulf of Mexico. J. Struct. Geol. 19, 463–480. Tavani, S., Storti, F., 2011. Layer-parallel shortening templates associated with double-
Rowan, M.G., Peel, F.J., Vendeville, B.C., 2004. Gravity-driven foldbelts on passive margins. edge fault-propagation folding. In: McClay, K., Shaw, J.H., Suppe, J. (Eds.), Thrust-
In: McClay, K.R. (Ed.), Thrust Tectonics and Hydrocarbon Systems AAPG Mem. 82, fault Related Folding AAPG Mem. 94, 121–135.
157–182. Tavani, S., Storti, F., Salvini, F., 2005. Rounding hinges to fault-bend folding: geometric and
Salvini, F., Storti, F., 2001. The distribution of deformation in parallel fault-related folds kinematic implications. J. Struct. Geol. 27, 3–22.
with migrating axial surfaces: comparison between fault-propagation and fault- Tavani, S., Storti, F., Salvini, F., 2006. Double-edge fault-propagation folding: geometry and
bend folding. J. Struct. Geol. 23, 25–32. kinematics. J. Struct. Geol. 28, 19–35.
Salvini, F., Storti, F., 2004. Active-hinge-folding-related deformation and its role in hydro- Tavani, S., Storti, F., Salvini, F., 2007. Modelling growth stratal architectures associated
carbon exploration and development — insights from HCA modeling. In: McClay, K.R. with double edge fault-propagation folding. Sediment. Geol. 196, 119–132.
(Ed.), Thrust tectonics and hydrocarbon systems AAPG Mem. 82, 453–472. Tavarnelli, E., 1997. Structural evolution of a foreland fold-and-thrust belt: the Umbria–
Sanderson, D.J., 1982. Models of strain variation in nappes and thrust sheets: a review. Marche Apennines, Italy. J. Struct. Geol. 19, 523–534.
Tectonophysics 88, 201–233. Thomas, G.S.P., Chiverrell, R.C., 2007. Structural and depositional evidence for repeated
Savage, H.M., Cooke, M.L., 2003. Can flat–ramp–flat fault geometry be inferred from fold- ice-marginal oscillation along the eastern margin of the Late Devensian Irish Sea Ice
shape?: A comparison of kinematic and mechanical folds. J. Struct. Geol. 25, 2023–2034. Stream. Quat. Sci. Rev. 26, 2375–2405.
Scharer, K.M., Burbank, D.W., Chen, J., Weldon, R.J., Rubin, C., Zhao, R., Shen, J., 2004. Thorbjornsen, K.L., Dunne, W.M., 1997. Origin of a thrust-related fold: geometric vs
Detachment folding in the Southwestern Tian Shan-Tarim foreland, China: shortening kinematic tests. J. Struct. Geol. 19, 303–319.
estimates and rates. J. Struct. Geol. 26, 2119–2137. Tindall, S.E., Davis, G.H., 1999. Monocline development by oblique-slip fault-propagation
Sharp, I.R., Gawthorpe, R.L., Underhill, J.R., Gupta, S., 2000. Fault-propagation folding in folding: the East Kaibab monocline, Colorado Plateau, Utah. J. Struct. Geol. 21,
extensional settings: examples of structural style and synrift sedimentary response 1303–1320.
from the Suez rift, Sinai, Egypt. GSA Bull. 112, 1877–1899. Tozer, R.S.J., Butler, R.W.H., Corrado, S., 2002. Comparing thin- and thick-skinned tectonic
Shaw, J.H., Shearer, P.M., 1999. An elusive blind-thrust beneath metropolitan Los Angeles. models of the Central Apennines, Italy. EGU Stephan Mueller Spec. Publ. Ser. 1,
Science 283, 1516–1518. 181–194.
Shaw, J.H., Suppe, J., 1994. Active faulting and growth folding in the eastern Santa Barbara Turner, J.P., Williams, G.A., 2004. Sedimentary basin inversion and intra-plate shortening.
Channel, California. GSA Bull. 106, 607–626. Earth-Sci. Rev. 65, 277–304.
Shaw, J.H., Suppe, J., 1996. Earthquake hazards of active blind-thrust faults under the Twiss, R.J., Moores, E.M., 1992. Structural Geology. Freeman and Company, New York
central Los Angeles basin, California. J. Geophys. Res. 101, 8623–8642. (532 pp.).
Shaw, J.H., Plesch, A., Dolan, J.F., Pratt, T.L., Fiore, P., 2002. Puente Hills blind-thrust system, Vann, I.R., Graham, R.H., Hayward, A.B., 1986. The structure of mountain fronts. J. Struct.
Los Angeles, California. Bull. Seismol. Soc. Am. 92, 2946–2960. Geol. 8, 215–227.
Shaw, J.H., Connors, C., Suppe, J., 2005. Seismic interpretation of contractional fault- Vergés, J., Marzo, M., Munoz, J.A., 2002. Growth strata in foreland settings. Sediment. Geol.
related folds: an AAPG seismic atlas. AAPG Stud. Geol. 53. 146, 1–9.
Smart, K.J., Ferrill, D.A., Morris, A.P., McGinnis, R.N., 2012. Geomechanical modelling of Waltham, D., 1989. Finite difference modeling of hangingwall deformation. J. Struct. Geol.
stress and strain evolution during contractional fault-related folding. Tectonophysics 11, 433–437.
576–577, 171–196. Waltham, D., Hardy, S., 1995. The velocity description of deformation. Paper 1: theory.
Song, T., Cawood, P.A., 2001. Effects of subsidiary faults on the geometric construction of Mar. Pet. Geol. 12, 153–163.
listric normal fault systems. AAPG Bull. 85, 221–232. Wilkerson, M.S., Medwedeff, D.A., Marshak, S., 1991. Geometrical modeling of fault-
Spang, J.H., McConnell, D.A., 1997. Effect of initial fault geometry on the development of related folds: a pseudo-three-dimensional approach. J. Struct. Geol. 13, 801–812.
fixed-hinge, fault-propagation folds. J. Struct. Geol. 19, 1537–1541. Williams, R.T., 1987. Energy balance for large thrust sheets and fault-bend folds. J. Struct.
Stewart, S.A., 1996. Influence of detachment layer thickness on style of thin-skinned Geol. 9, 375–379.
shortening. J. Struct. Geol. 18, 1271–1274. Williams, G., Chapman, T., 1983. Strains developed in the hangingwalls of thrusts due to
Stille, H., 1924. Grundfragen der vergleichenden Tektonik. Verlag von Gebrüder their slip/propagation rate: a dislocation model. J. Struct. Geol. 5, 563–571.
Bornträger, Berlin (443 pp.). Williams, G.D., Powell, C.M., Cooper, M.A., 1989. Geometry and kinematics of inversion
Stockmal, G.S., Lebel, D., McMechan, M.E., MacKay, P.A., 2001. Structural style and evolu- tectonics. In: Cooper, M.A., Williams, G.D. (Eds.), Inversion Tectonics Geol. Soc. Spec.
tion of the triangle zone and external foothills, southwestern Alberta: implications for Publ. 44, 3–15 (London).
thin-skinned thrust-and-fold belt mechanics. Bull. Can. Petrol. Geol. 49, 472–496. Willis, B., 1894. The Mechanics of Appalachian Structure. Department of the Interior —
Storti, F., Poblet, J., 1997. Growth stratal architectures associated to decollement folds and U.S. Geological Survey (281 pp.).
fault-propagation folds. Inferences on fold kinematics. Tectonophysics 282, 353–373. Wiltschko, D.V., 1979. A mechanical model for thrust sheet deformation at a ramp.
Storti, F., Salvini, F., 1996. Progressive rollover fault-propagation folding: a possible kine- J. Geophys. Res. 84, 1091–1104.
matic mechanism to generate regional-scale recumbent folds in shallow foreland Wiltschko, D.V., Groshong Jr., R.H., 2012. The Chamberlin 1910 balanced section: context,
belts. AAPG Bull. 80, 174–193. contribution, and critical reassessment. J. Struct. Geol. 41, 7–23.
Storti, F., Salvini, F., McClay, K., 1997. Fault-related folding in sandbox analogue models of Withjack, M.O., Olson, J., Peterson, E., 1990. Experimental models of extensional forced
thrust wedges. J. Struct. Geol. 19, 583–602. folds. AAPG Bull. 74, 1038–1054.
Strayer, L.M., Hudleston, P.J., 1997. Numerical modeling of fold initiation at thrust ramps. Woodward, N.B., 1997. Low-amplitude evolution of break-thrust folding. J. Struct. Geol.
J. Struct. Geol. 19, 551–566. 19, 293–301.
Strayer, L.M., Erickson, S.G., Suppe, J., 2004. Influence of growth strata on the evolution of Wu, S., Yu, Z., Zhang, R., Han, W., Zou, D., 2005. Mesozoic–Cenozoic tectonic evolution of
fault-related folds — distinct-element models. In: McClay, K.R. (Ed.), Thrust tectonics the Zhuanghai area, Bohai-Bay Basin, east China: the application of balanced cross-
and hydrocarbon systemsAAPG Mem. 82, 413–437. sections. J. Geophys. Eng. 2, 158–168.
Stüwe, K., 2000. Geodynamik der Lithosphäre. Springer Verlag, Berlin (405 pp.). Yamada, Y., McClay, K., 2004. 3-D Analog modeling of inversion thrust structures.
Suppe, J., 1983. Geometry and kinematics of fault-bend folding. Am. J. Sci. 283, 684–721. In: McClay, K.R. (Ed.), Thrust tectonics and hydrocarbon systemsAAPG Mem.
Suppe, J., Medwedeff, D.A., 1990. Geometry and kinematics of fault-propagation folding. 82, 276–301.
Eclogae Geol. Helv. 83, 409–454. Yeats, R.S., Huftile, G.J., 1995. The Oak Ridge fault system and the 1994 Northridge
Suppe, J., Chou, G.T., Hook, S.C., 1992. Rates of folding and faulting determined from earthquake. Nature 373, 418–420.
growth strata. In: McClay, K. (Ed.), Thrust Tectonics. Chapman and Hall, London, Yielding, G., Needham, T., Jones, H., 1996. Sampling of fault populations using sub-surface
pp. 105–121. data: a review. J. Struct. Geol. 18, 135–146.
Suppe, J., Sàbat, F., Muñoz, J.A., Poblet, J., Roca, E., Vergés, J., 1997. Bed-by-bed fold growth Zehnder, A.T., Allmendinger, R.W., 2000. Velocity field for the trishear model. J. Struct.
by kink-band migration: Sant Llorenç de Morunys, eastern Pyrenees. J. Struct. Geol. Geol. 22, 1009–1014.
19, 443–461. Ziesch, J., Tanner, D.C., Krawczyk, C.M., 2014. Strain associated with the fault-parallel flow
Suppe, J., Connors, C.D., Zhang, Y., 2004. Shear fault-bend folding. In: McClay, K.R. (Ed.), algorithm during kinematic fault displacement. Math. Geosci. 46 (1), 59–73.
Thrust Tectonics and Hydrocarbon Systems AAPG Mem. 82, 303–323. Zoetemeijer, R., Sassi, W., 1992. 2-D reconstruction of thrust evolution using the fault-
Tanner, D.C., Behrmann, J.H., Dresmann, H., 2003. Three-dimensional retro-deformation bend fold method. In: McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall,
of the Lechtal Nappe, Northern Calcareous Alps. J. Struct. Geol. 25, 737–748. London, pp. 133–140.
Tanner, D.C., Brandes, C., Leiss, B., 2010. Structure and kinematics of an outcrop-scale fold- Zoetemeijer, R., Sassi, W., Roure, F., Cloetingh, S., 1992. Stratigraphic and kinematic
cored triangle zone. AAPG Bull. 94, 1799–1809. modeling of thrust evolution, northern Apennines, Italy. Geology 20, 1035–1038.

View publication stats

You might also like