Deciphering Thrust Fault Nucleation and Propagation and The

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 12

Journal of Structural Geology 85 (2016) 1e 11

Contents lists available at ScienceDirec t

Journal of Structural Geology


journal homepage: www.elsevier.com/locate/js g

Deciphering thrust fault nucleation and propagation and the


importance of footwall synclines

David A. Ferrill a, *, Alan P. Morris a, Sarah S. Wigginton a, Kevin J. Smart a, Ronald N.


McGinnis a, Daniel Lehrmann b
a
Department of Earth, Material and Planetary Sciences, Southwest Research Institute ®, 6220 Culebra Road, San Antonio, TX 78238, USA b Geoscience
Department, Trinity University, One Trinity Place, San Antonio, TX 78212, USA

articleinfo abstract

Article history: In this paper, we analyze small scale examples of thrust faults and related folding in outcrops
Received 27 October 2015 of the Cretaceous Boquillas Formation within Big Bend National Park in west Texas to
Received in revised form develop detailed understanding of the fault nucleation and propagation that may aid in the
15 January 2016 interpretation of larger thrust system structure. Thrust faults in the outcrop have maximum
Accepted 27 January 2016 displacements ranging from 0.5 cm to 9 cm within competent limestone beds, and these
Available online 10 displacements diminish both upward into anticlines and downward into synclines within the
February 2016
interbedded and weaker mudrock layers. We interpret the faults as having nucleated within the
competent units and partially propagated into the less competent units without developing
Keywords: floor or roof thrusts. Faults that continued to propagate resulted in hanging wall anticlines
Thrust fault above upwardly propagating fault tips, and footwall synclines beneath downwardly
Nucleation propagating fault tips. The observed structural style may provide insights in the nucleation of
Footwall syncline faults at the formation scale and the structural development at the mountain-range scale.
Boquillas formation Decollement or detach- ment layers may be a consequence rather than cause of thrust ramps
Eagle Ford Formation through competent units and could be over interpreted from seismic data.
© 2016 Elsevier Ltd. All rights reserved.
1. Introduction
Armstrong and Bartley, 1993; Wickham, 1995; Hughes et
Thrust faults are often observed to “ramp” through al., 2014; Hughes and Shaw, 2014, 2015), and footwall
competent strata from a basal detachment, forming as synclines are sometimes recognized and often interpreted
splays or branches from a regional low-angle thrust or as forming due to frictional drag or fault propagation
detachment in a weak or incompetent stratum (e.g., Rich, through the limb or along the axial surface of an anticline
1934; Rodgers, 1950; Dahlstrom, 1970; Boyer and Elliot,
(e.g., Link, 1949; Gallup, 1951; Armstrong, 1968; Fleck,
1982; Butler, 1982; Williams and Chapman, 1983; Mitra,
1970; Perry, 1978; Kulander and Dean, 1986; Fischer et
1992; Allmendinger et al., 2004; Hardy and Allmendinger,
al., 1992; McNaught and Mitra, 1993; Thorbjornsen and
2011; McClay, 2011; Hughes et al., 2014; Hughes and
Dunne, 1997; Wallace and Homza, 2004). However, some
Shaw, 2014, 2015). The basal low-angle thrust is often
workers have suggested that thrust faults nucleate within a
explicitly or tacitly assumed to form first and other faults to
stratigraphic package and propagate upward and downward
emerge or grow from it, ramping upward to higher
rather than ramping up from a detachment horizon (e.g.,
stratigraphic levels, although this is not necessarily the case
Chapman and Williams,1984; Eisenstadt and DePaor,1987;
and ramps may form first and flats later (e.g., Chapple,
Apotria and Wilkerson, 2002; Tavani et al., 2006; Uzkeda
1978). Hanging wall anticlines are common and typically
et al., 2010), and some footwall synclines have been
interpreted as fault-bend or faultpropagation folds (e.g.,
interpreted to be formed by downward thrust fault-tip
Suppe, 1983; Jamison, 1987; Chester and Chester, 1990;
propagation (Williams and Chapman, 1983; Ramsay,
Suppe and Medwedeff, 1990; Mitra, 1992;
1992; Morley, 1994; McConnell et al., 1997; Welch et al.,
2009; Uzkeda et al., 2010). Eisenstadt and DePaor (1987)
* Corresponding author. E-mail address: [email protected] (D.A. Ferrill).
proposed an alternative general model for thrust fault system development wherein the thrust ramps nucleate first within the
competent units, and displacement in incompetent layers subsequently develops to maintain displacement compatibility in
the
http://dx.doi.org/10.1016/j.jsg.2016.01.009
2 D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12
0191-8141/© 2016 Elsevier Ltd. All rights reserved.
system. wall anticlines form through this process can help to
In this paper, we analyze small scale examples of thrust identify the location of fault nucleation within a
faults and related folding in outcrops of the Cretaceous mechanically stratified section. Understanding this process
Boquillas Formation within Big Bend National Park in and the mechanical stratigraphy of a particular area can be
west Texas (Figs. 1 and 2) to develop detailed used to predict fault-related folds, including footwall
understanding of the fault nucleation and propagation that synclines which may be difficult to image with seismic
may aid in the interpretation of larger thrust system reflection data (e.g., Apotria and Wilkerson, 2002).
structure. Our observations indicate that faults that ramp
through competent layers likely nucleate in those
2. Geologic setting
competent layers and propagate upward and downward
rather than forming as splay faults propagating from a
2.1. Tectonic setting
layer-parallel or low-angle detachment. Maximum thrust
fault displacements are localized in the competent
The study area is a location known as Ernst Tinaja in
carbonate (limestone or chalk) layers, and they lose
Big Bend National Park in Brewster County, west Texas,
displacement into mudrock layers, as illustrated in the
USA (Fig. 1). Ernst Tinaja is along a gorge and arroyo
distance vs. displacement diagram in Fig. 3. Where fault
incised through Cuesta Carlota e a ridge and tilted fault
propagation is impeded by incompetent strata, fault tip
block of west-dipping Cretaceous strata in the footwall of
folding occurs, producing hanging wall anticlines
an east-dipping normal fault (Fig. 2; Maxwell et al., 1967;
associated with the upwardly propagating fault tips and
Moustafa, 1988). The arroyo drains the Ernst Basin half
footwall synclines associated with downwardly
graben east of the normal fault. Cuesta Carlota lies along
propagating fault tips e similar to the ‘thrust-tip folds’ of
the western edge of the Sierra Del Carmen, which
Chapman and Williams (1984, 1985). This fault tip folding
experienced northeast-directed Laramide contraction
is caused by the combination of fault slip within the
(Maxwell et al., 1967; Moustafa, 1988; Maler, 1990;
competent layers and the ability of incompetent layers to
Lehman, 1991; Turner et al., 2011), followed by
absorb displacement by folding. Such folds are therefore
southwest-directed Basin and Range extension (Maxwell et
the result of fault arrest and related fault propagation
al.,1967; Moustafa,1988; Turner et al., 2011). The faults
folding. Recognizing that footwall synclines and hanging

Fig. 1. Shaded relief map showing the Sierra del Carmen and the location of Figure 2. Inset map shows location with respect to Texas. Northwestern
limit of macroscale Paleozoic Ouachita deformation (“Ouachita Orogenic Front”) is after Flawn et al. (1961). Northeastern limit of Laramide
macroscale deformation (“Laramide Orogenic Front”) is based on a compilation of regional mapping and interpretations of Erdlac (1994).
Northeastern limit of Basin and Range extensional deformation (“Basin and Range Extension”) is after Stewart et al. (1998).
D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e11 3

Fig. 2. Geologic map showing location of the Ernst Tinaja study site, Cuesta Carlota, and Ernst Basin (after Turner et al., 2011). Inset stratigraphic
section is based on thicknesses from measured sections #19 (Aguja and Pen Formations), #21 (Ernst and San Vicente Members of the Boquillas
Formation and the Buda Formation), and #2 (Del Rio Clay and Santa Elena Limestone) of Maxwell et al. (1967). *Quaternary and Tertiary
stratigraphic sections are bounded by unconformities and their thicknesses are highly variable in the map area.
highlighted limestone bed. The central portion of this bed (location of
displacement maximum) is the inferred location where the shear failure
initiated. The center of the green-highlighted limestone bed is 3.0 m
above the top of the Buda Limestone.
under investigation here formed during regional Laramide
contractional deformation and were subsequently cut by
Basin and Range extensional structures, in this case,
calcite- filled extension fractures (veins).

2.2. Stratigraphy

The stratigraphic section at Ernst Tinaja includes a


massive carbonate section of the Santa Elena Limestone, a
very thin Del Rio Clay section, the massive Buda
Limestone, the complete Ernst Member of the Boquillas
Formation (equivalent of the Eagle Ford Formation;
Maxwell et al., 1967), and the San Vicente Member of the
Fig. 3. Thrust fault nucleated in limestone bed propagated upward and
downward into less competent mudrock beds. Arrested propagation with Boquillas Formation (equivalent of the Austin Chalk;
continued displacement caused fault tip folding and produced the Maxwell et al., 1967, Fig. 2). Our focus in this paper is on
footwall syncline and hanging wall anticline. Outcrop location is at limestone and mudrock beds in the upper part of the Ernst
Ernst Tinaja in Big Bend National Park, Texas (USA). Note that Member of the
bedding adjacent to the fault between the intersection of the hanging
wall anticline axial surface and the footwall syncline axial surface with
Boquillas Formation.
the fault is not folded. Inset displacement versus distance diagram for The Ernst Member of the Boquillas Formation is the
the fault illustrates maximum displacement of 53 cm in the green equivalent of the Eagle Ford Formation in central and south
4 D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12
Texas (Maxwell et al., 1967), which has recently been This section is a marine, predominantly pelagic succession
developed as a major unconventional oil and gas reservoir of calcareous mudrock, nannoplankton and pelagic
(e.g., Carrillo, 2010; Donovan and Staerker, 2010; Lock et foraminifer microgranular packstone (chalk), thin beds of
al., 2010; Hentz and Ruppel, 2010; Treadgold et al., 2010; intercalated
Basu et al., 2012; Bodziak et al., 2014; Ferrill et al., 2014).
Fig. 4. (a) Uninterpreted outcrop photomosaic. Outcrop is 10 m long. (b) Outcrop photomosaic annotated with structural features and measurement
locations, with competent units C1, C2, C3, and C4 color coded. Insets show stereonets of structural data, lithostratigraphic section, and Schmidt
hammer rebound profile. Dots on great-circle fault traces on stereonets are measured slickenline orientations.
6 D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12
Fig. 5. Examples of thrust faults tipping upward and downward, and associated folds. (a) Southwest-vergent thrust fault with folding at upper and
lower fault tips indicating propagation upward and downward out of competent bed C3. (b) Northeast-vergent thrust fault with folding at upper and
lower fault tips indicating propagation upward and downward out of competent bed C3. (c) Three southwest-vergent thrust faults each tip into folds
at upper and lower fault tips indicating propagation upward and downward out of competent beds C2 and C3. Note thinning of incompetent bed
between adjacent footwall syncline beneath f6 and hanging wall anticline of f32. Displacement versus distance diagrams illustrate trapezoidal
displacement profiles, constrained by the measured displacement through the limestone beds and abrupt tips in the mudrock beds. The horizontal
length of the shaded area under the curve represents the faulted distance through the limestone bed. Orange multi-tool is 8.1 cm long.
stratification, fragmented skeletal material and are
winnowed indicating storm agitation on the seafloor.
Because of the abrupt changes in hydrodynamic regime
associated with storm events, the bedding boundaries
within the hydrodynamic packstone-grainstone and
heterolithic facies are sharp. The overlying San Vicente
Member of the Boquillas Formation (Austin Chalk
equivalent) is composed of calcareous mudrock and chalk

Fig. 6. Field photograph of stepped crystal-fiber slickensides.

Fig. 8. Northwest-striking bed-perpendicular calcite veins cut low-angle


thrust faults, (a) fault 31, and (b) fault 16, indicating early contractional
deformation in thrust faulting stress regime followed by extension
fracturing. Orange multi-tool is 8.1 cm long. See Fig. 4b for the
locations of these faults.

Fig. 7. Graph of fault displacements versus thickness of competent units intercalations similar to the pelagic portions of the Ernst
2, 3, and 4. Member except that the chalk layers exhibit greater
bioturbation and contain benthic fauna indicating more
oxygenation of the bottom waters. The section that is the
focus of the present study is within the uppermost 3 m of
calcareous mudrock and limestone layers, with intervals of
the Ernst Member, composed of alternating hydrodynamic
hydrodynamically agitated skeletal and planktonic
lenticular packstone-grainstone beds and mudrock beds
foraminifer lime grainstone-packstone beds and numerous
with relatively sharp lithologic contacts (Fig. 4a, b).
thin volcanic ash layers.
Much of the Ernst Member consists of a cyclic
alternation of calcareous mudrock and chalk with both
distinct and gradational bedding contacts. Hydrodynamic
packstone-grainstone beds are lenticular to continuous,
contain ripple cross-lamination, hummocky cross
2.3. Structural style

Deformation features in the Ernst Member of the


Boquillas Formation at Ernst Tinaja include contractional
features that formed in response to northeast-directed
contraction: smalldisplacement thrust faults, contractional
folds, and tectonic stylolites (e.g., Figs. 3e7). In addition,
two sets of bed-perpendicular systematic opening-mode
extension fractures (joints and veins) are present including
(1) a dominant northwest-striking, often calcite-filled, set
(several are represented in our data; Figs. 4b and 8), and (2)
an orthogonal northeast striking set (primarily joints) that
are subsidiary to and generally abut against the northwest
striking set (best seen in nearby pavements and not
measured in this study). The dominant opening-mode
extension fracture set (northwest-striking calcite veins) cuts
the thrust faults (Fig. 8). In some places the opening mode
veins reactivate tectonic stylolites, forming coplanar with
but cutting the teeth of tectonic stylolites, indicating that
the southwest directed extension that
8 D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12

Table 1
Data measured from exposure of contractional faults along the Ernst Tinaja arroyo. Fault and vein orientations are reported in both present-day orientations (unrotated) and corrected for bedding
strike/dip of 136/21 (rotated).* “Fault Spacing” is the distance from previous fault measurement to the SW measured in the shortening direction. Values in parentheses for f1 and f14 are minimum
distances, measured as the unfaulted distance from the end of the outcrop to the fault. f# SS indicates strike slip fault orientation. C2, C3, and C4 refer to competent units 2, 3, and 4 respectively e see
Fig. 4b for details. Strike, dip, and rake measurements are reported following the right-hand rule where the dip direction is to the right of the reported strike azimuth. “Rake” denotes the angle of the
slickenlines on the fault surface and was measured clockwise from the reported strike azimuth.

Number Unit Disp. (cm) *Fault Unrotated Rotated (136/21)


Spacing (cm)
Strike Dip Rake Strike Dip Rake

Thrust Faults
f1 C3 2 (17) 115 21 85 36 7 166
f2 C3 3 29 141 10 90 311 11 100
f3 C3 0.5 75 148 48 90 155 28 81
f4 C3 4 79 130 31 95 119 10 107
f5 C3 5 103 176 49 94 194 35 70
f6 C3 6 56 308 11 95 313 32 90
f7ss C3 <1 90 238 88 10 58 88 11
f8 C3 <5 4 215 5 178 302 21 90
f9 C3 1.2 71 350 15 90 331 34 111
f10 C3 5 25 54 24 18 12 29 65
f11 C3 5 55 335 4 98 319 25 114
f12 C3 4.5 73 166 50 90 181 33 70
f13 C3 5 56 8 5 80 326 24 123
f14 C4 3 (39) 237 26 e 269 36 e
f15 C4 4.5 60 184 35 e 237 19 e
f16 C4 9 100 145 41 90 153 20 81
f17 C4 1.5 94 174 50 90 191 35 67
f18 C4 2 141 353 13 72 331 32 96
f19 C4 <1 22 223 10 e 289 23 e
f20 C4 6 16 235 12 0 288 26 125
f21 C4 5.5 160 151 30 93 179 11 63
f22 C4 2 12 90 27 110 42 19 161
f23 C4 4 44 328 0 50 316 21 62
f24 C4 0.5 58 180 5 80 305 18 135
f25ss C4 1.5 152 84 80 170 79 68 8
f26 C4 2.5 41 15 3 160 323 23 32
f27 C4 2 15 235 5 164 303 22 95
f28 C4 <1 22 160 28 e 202 12 e
f29 C4 1.5 36 40 12 90 345 25 148
f30 C2 4 e 151 34 90 171 15 69
f31 C2 7.5 e 20 10 90 336 27 136
f32 C2 5 e 215 2 90 310 21 175
f33 C2 6.5 e 160 49 80 173 31 64
f34 C2 1 e 150 60 90 155 40 82
Veins
v1 C3 e e 349 81 e 169 81 e
v2 C3 e e 0 74 e 358 89 e
v3 C3 e e 358 84 e 178 80 e
v4 C3 e e 353 85 e 174 78 e
v5 C3 e e 353 80 e 173 83 e
v6 C3 e e 355 78 e 174 86 e
v7 C3 e e 350 80 e 170 82 e
v8 C3 e e 341 75 e 160 86 e
v9 C4 e e 347 81 e 167 81 e
v10 C4 e e 340 75 e 159 86 e
v11 C4 e e 344 77 e 163 84 e
v12 C4 e e 352 77 e 171 86 e
v13 C4 e e 350 70 e 348 88 e
v14 C4 e e 352 78 e 171 85 e
v15 C4 e e 348 87 e 169 75 e
v16 C4 e e 338 77 e 158 83 e
v17 C4 e e 331 71 e 150 89 e
v18 C2 e e 350 73 e 168 89 e
Bedding
b1 C4 e e 143 23 e e e e
b2 C3 e e 135 24 e e e e
b3 C4 e e 131 17 e e e e

formed these structures post-dated the northeast directed contraction. With fractures forming in response to regional Basin and Range extension (25-2
respect to the well-established regional tectonic framework, these observed Ma according to Turner et al., 2011).
structural elements and cross cutting relationships are consistent with 3. Analysis of thrust faults in the Boquillas Formation at Ernst Tinaja, Big
formation of Late Cretaceous-early Tertiary (70-50 Ma; Lehman, 1991) Bend National Park
Laramide structures that formed in response to northeast-directed
contraction and are overprinted by the northwest-striking opening-mode
D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12 9

3.1. Mechanical stratigraphy Thirty-two of the faults have measured thrust displacements in the
range from 0.5 cm to 9 cm. The 13 faults measured in C3 have
We used an N-type Schmidt rebound hammer to characterize the displacements ranging from 0.5 to 6 cm, average displacement of 3.6 cm,
mechanical stratigraphy for the 1.6 m section of stratigraphy in this study. The and average spacing of 56 cm. The 16 faults measured in C4 have
section was subdivided into 10 mechanical layers ranging in thickness from 9 displacements ranging from 0.5 to 9 cm, average displacement of 3.0 cm,
cm to 30 cm. Each mechanical unit was tested 10 times (near the centerline of and average spacing of 63 cm. Maximum displacements are on the order of
the bed within a bed parallel distance of 25 cm) using the Schmidt hammer, half the competent unit thickness (Fig. 7). Two strike-slip faults were also
and the 10 values were averaged and used to plot the rebound profile shown in measured e one from C3 and one from C4 e with displacements of <1 cm
Fig. 4b. The section contains four competent (high rebound) limestone- and 1.5 cm, respectively. Tip lines of these strike-slip faults were not
dominated layers from the bottom up that we refer to as competent unit 1 observable in the outcrop. Total shortening by the faults measured in C3
(C1),11.5 cm thick; competent unit 2 (C2),16.5 cm thick; competent unit 3 sums to 5.1%, and in bed C4 sums to 3.7%.
(C3), 9 cm thick; and competent unit 4 (C4), 18.5 cm thick (Fig. 4b). The thrust faults all have relatively uniform displacement along their
sections through the central part of competent units, and each of the faults
3.2. Thrust faults tips into mudrock beds above and below (Fig. 5). Inset displacement versus
distance diagrams in Fig. 5 illustrate trapezoidal displacement profiles,
In the exposure, we measured 34 faults within mechanical units C2, C3, constrained by the measured (relatively uniform) displacement through the
and C4 (Fig. 4b; Table 1) along mid-bed scanline surveys. Units C3 and C4 limestone beds and abrupt displacement gradients at the fault tips in the
are exposed over lateral bed-parallel distances of 8 m and 10.5 m, bounding mudrock beds. Folds are consistently developed at these upward
respectively, measured in the dominant fault slip (slickenline) direction. We and downward fault tips. Folding in the weak layers is synchronous with
measured a total of 13 faults in C3, and 16 faults in C4, and these measured faulting in the strong layers. Where faults have propagated through folds in
faults constitute every recognized fault that encountered the mid-lines of these the weaker layers, hanging wall anticlines have developed above faults at
two limestone beds. In addition, 5 faults were measured from C2. Although upward tips, and footwall synclines below faults at downward tips. Careful
other faults are present in C2, they were not measured because of outcrop inspection reveals no evidence of discrete faults, sliding surfaces, or shear
quality and safety concerns. The studied thrust faults constitute a conjugate zones within the mudrock layers parallel to layering. Contractional folding
array with vergence to both the northeast and southwest (Fig. 4b). All of the and thickness changes are evident in the vicinity of fault tips within the
measured faults in the array are consistent with internal layer parallel mudrock layers.
shortening during Laramide contractional deformation. Correcting for average
bedding strike of 136 and dip of 21, the contractional faults rotate back to two 4. Discussion
conjugate fault sets with an average fault dip of 24 (Table 1; Fig. 4b).
The faults are all marked with stepped crystal-fiber slickensides (Fig. 6; Schmidt hammer Rebound (R) data measured in outcrop indicate
c.f., Hancock, 1985), and are composed of both shear and dilational segments. different mechanical behavior for mudrock versus limestone. Although
Calcite fibers link across dilational segments, indicating continuous growth as these measurements are likely altered from original character e particularly
fault slip accumulated (Ramsay and Huber, 1983) e a similar style to small- more clay rich mudrock beds e they suggest mechanical differences
displacement normal faults in the Eagle Ford Formation described by Ferrill between these contrasting rock types. R is frequently used as a proxy for
et al. (2014). Two faults display stylolitic suturing indicating pressure solution the Young's modulus (as well as unconfined compressive strength; Katz et
at restraining bends between fault segments. al., 2000; Aydin and Basu, 2005). Strain compatibility implies that layer-
parallel shortening is (almost) identical in the different layers of Fig. 4. At
very small strain values, prior to the initiation of faulting and when
LeMoigne thrust and the associated footwall syncline axial surface. Vertical relief in the
photograph is 1000 m.
elastic behavior would be a valid approximation, constant strain and different
Young's moduli imply different layer-parallel stresses in the different layers.
Higher layer-parallel stresses would be expected (according to elasticity
theory) in layers with higher Young's modulus, i.e. with higher R values. The
overburden stress (vertical load) would have been essentially uniform for all
beds analyzed. Thus differential stresses would have been greater in the layers
with higher R values, and it was in these where faulting was initiated.
Recognizing the structural style discussed here e in particular asymmetric
hanging wall anticlines associated with upwardly propagating fault tips and
asymmetric footwall synclines associated with downwardly propagating fault
tips e can be used to identify the nucleation point or zone for thrust faults. By
analogy, it is likely that some large thrust ramps nucleate in massive
competent units of low ductility that lack incompetent interbeds to arrest the
propagation of the fault, dictated by the large scale mechanical stratigraphy
and crustal stress fields extant at the time of deformation.
Fig. 9. Field photograph of the footwall syncline beneath the LeMoigne thrust, exposed in the The term fault-propagation fold is commonly used to describe folds
footwall of the Panamint fault in eastern California. Photo is annotated with the traces of the formed above upwardly propagating thrust faults. In this situation of an
arrested fault tip, continued fault displacement is
10 D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12

Fig. 10. Footwall synclines beneath the Wheeler Pass Thrust and Lee Canyon Thrust in the Spring Mountains in southern Nevada, from Burchfiel et al. (1974). No vertical exaggeration.

Fig. 11. Schematic illustration for fault nucleation and propagation in (a) contractional, and (b) and (c) extensional structural regimes. Beds in the fault nucleation layers are not folded (dark gray in a,
grey in b and c), whereas incompetent or mechanically layered units at and beyond an arrested fault tip are folded prior to the fault propagating through. In the case of a thrust fault (a), this produces a
hanging wall anticline above the upwardly propagating fault tip and a footwall syncline below the downwardly propagating fault tip. Conversely, in the case of a normal fault (b and c), this process
produces a hanging wall syncline above the upwardly propagating fault tip and a footwall anticline below the downwardly propagating fault tip. (b) displays a case for a normal fault terminated in
incompetent layers, and (c) shows the normal fault after it has propagated through the incompetent layers.
accommodated by folding within incompetent or mechanically layered strata footwall synclines is the product of the same process of folding near and
beyond the fault tip (e.g., Williams and Chapman, 1983). The formation of beyond an arrested lower fault tip during continued fault displacement.
D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12 11

The processes inferred here are dependent on mechanical stratigraphy and National Park (west Texas, USA) provides well-exposed examples of
stress regime, and are likely to influence contractional deformation style at all contractional structures that formed during regional Laramide contractional
scales. Footwall synclines beneath thrust faults occur at macroscale. For deformation and were subsequently cut by Basin and Range extensional
example, the footwall of the Permian age LeMoigne thrust in eastern structures (calcite veins). Thrust faults in the outcrop have maximum
California (Snow, 1992) provides a clear example of a footwall syncline. displacements ranging from 0.5 cm to 9 cm within competent limestone
Massive Cambrian and Ordovician dolomite layers are thrust over thinly beds, and these displacements diminish both upward into anticlines and
bedded limestone, shale, and siltstone of the Keeler Canyon Formation, which downward into synclines within the interbedded and weaker mudrock
is folded into a footwall syncline (Fig. 9; Hall, 1971). This footwall syncline layers. These faults nucleated within the competent units and partially
and the LeMoigne thrust are exposed in the footwall of e and cut by e the propagated into the less competent units without developing floor or roof
active Panamint Valley normal fault (Hall, 1971; Wernicke et al., 1988; Snow thrusts. Faults that continued to propagate resulted in hanging wall
and Wernicke, 1989; Snow, 1992). anticlines above upwardly propagating fault tips, and footwall synclines
Other large-scale examples of footwall synclines are welldocumented in beneath downwardly propagating fault tips. The structural style is
the western Basin and Range region (Burchfiel, 1965; Burchfiel et al., 1974). analogous to that within mountain-range scale Cordilleran structures in the
The Sevier fold-thrust belt e part of the Rocky Mountain Cordillera that was Spring Mountains (Nevada, USA). We believe that the small-scale
subsequently dissected by Basin and Range normal and strike slip faulting e is structures described in this paper provide insights into the nucleation of
exposed in the Spring Mountains, Nevada. Detailed mapping and cross faults at the formation scale and structural development at the mountain
sections by Burchfiel (1965) and Burchfiel et al. (1974) document large-scale range scale. Structural styles of fold thrust belts have been studied
footwall synclines beneath thrust faults (c.f., Fig. 10). The Keystone thrust, intensively in recent decades, and low-angle floor and roof thrusts or
Lee Canyon thrust, and Wheeler Pass thrust each contain footwall synclines detachment horizons are often interpreted to provide kinematic fault
that have a vertical to overturned limb hundreds of meters to 1 km long. linkage through mechanically layered crustal sections. Based on the
Footwall syncline geometries such as those illustrated in Figs. 9 and 10 are outcrop results from our study at Ernst Tinaja, we deduce that decollement
often interpreted as forming due to frictional drag or fault propagation through or detachment faults are a consequence rather than a cause of thrust ramps
the limb or along the axial surface of an anticline (e.g., Link, 1949; Gallup, through competent units. If this structural style can be applied up to
1951; Armstrong, 1968; Fleck, 1970; Perry, 1978; Kulander and Dean, 1986; regional cross section scale, it is possible that widespread discrete floor and
Fischer et al., 1992; McNaught and Mitra, 1993; Thorbjornsen and Dunne, roof thrusts have been over interpreted.
1997; Wallace and Homza, 2004). However, the geometries in these examples
also match the structural style of the small scale structures documented in this Acknowledgments
paper, consistent with faults nucleating shallower in the crust and propagating
downward as illustrated in Fig. 11a and discussed by previous workers We thank Big Bend National Park for providing the research permit
(Williams and Chapman, 1983; Chapman and Williams, 1984; Eisenstadt and that allowed us to perform this research. Financial support for this work
DePaor, 1987; Ramsay, 1992; Morley, 1994; McConnell et al., 1997; Apotria was provided by Southwest Research Institute Internal Research and
and Wilkerson, 2002; Tavani et al., 2006; Welch et al., 2009; Uzkeda et al., Development project #R8588. We appreciate the constructive reviews by
2010). Large displacement on these thrusts has now placed a hanging wall Stefano Tavani, Nestor Cardozo, and Toru Takeshita that led to significant
lower flat on top of the footwall ramp and footwall syncline. This structural improvement of the manuscript.
evolution is in direct contrast with the generalized folding followed by
faulting model for these structures proposed by Fleck (1970). The thrust faults References
described in the Spring Mountains are cut later by high-angle normal faults
formed during Basin and Range extension (Fig. 10). Although it does not Allmendinger, R.W., Zapata, T., Manceda, R., Dzelalija, F., 2004. Trishear kinematic
modeling of structures, with examples from the Neuquen Basin, Argentina. In: McClay,
imply similarity of deformation mechanisms, it is interesting to note that this K.R. (Ed.), Thrust Tectonics and Hydrocarbon Systems, 82. AAPG Memoir, pp.
pattern is similar to that of the mesoscale structures at Ernst Tinaja where 356e371.
Laramide (Cordilleran) thrust faults are cut by subsequent extension veins that Apotria, T.G., Wilkerson, M.S., 2002. Seismic expression and kinematics of a fault related
we interpret to be associated with Basin and Range extension (Fig. 8). fold termination: Rosario structure, Maracaibo Basin, Venezuela. J. Struct. Geol. 24,
671e687.
These concepts for fault nucleation within competent beds, and both Armstrong, P.A., Bartley, J.M., 1993. Displacement and deformation associated with a
upward and downward fault propagation away from the nucleation site into lateral thrust termination, southern Golden Gate Range, southern Nevada, U.S.A. J.
the overlying and underling strata, closely parallels recent observations and Struct. Geol. 15, 721e735.
interpretations from extensional fault systems (Fig. 11). As normal faults Armstrong, R.L., 1968. Sevier orogenic belt in Nevada and Utah. Geol. Soc. Am. Bull. 79,
429e458.
propagate upwards out of a competent nucleation layer, fault tip folding Aydin, A., Basu, A., 2005. The Schmidt hammer in rock material characterization. Eng. Geol.
occurs in less competent beds as the fault tip is arrested (Ferrill et al., 2007, 81, 1e14.
2012). With continued propagation, the fault breaks along the footwall side of Basu, N., Barzola, G., Bello, H., Clarke, P., Viloria, O., 2012. Eagle Ford reservoir
characterization from multisource data integration. Am. Assoc. Pet. Geol. Search and Discovery
the monocline, and leaves the monocline attached to the hanging wall.
Article #80234.
Conversely, the downward propagation of the fault out of the competent Bodziak, R., Clemons, K., Stephens, A., Meek, R., 2014. The role of seismic attributes in
nucleation layer into less competent underlying beds produces a fault tip understanding the ‘frac-able’ limits and reservoir performance in shale reservoirs: an
monocline that eventually may be broken through and will generally be Example from the Eagle Ford Shale, South Texas, USA. AAPG Bull. 98, 2217e2235.
Boyer, S.E., Elliott, D., 1982. Thrust systems. AAPG Bull. 66, 1196e1230.
attached to the footwall side of the fault (Ferrill et al., 2011). In horizontal
Burchfiel, B.C., 1965. Structural geology of the Specter Range quadrangle, Nevada, and its
rocks, the hanging wall monocline will be a syncline, and the footwall regional significance. Geol. Soc. Am. Bull. 76, 175e192.
monocline will be an anticline e counterparts to the hanging wall anticline and Burchfiel, B.C., Fleck, R.J., Secor, D.T., Vincelette, R.R., Davis, G.A., 1974. Geology of
footwall syncline described in this paper associated with upward and the Spring Mountains, Nevada. Geol. Soc. Am. Bull. 85, 1013e1022.
downward thrust fault propagation from the competent layer in which the Butler, R.W.H., 1982. The terminology of structures in thrust belts. J. Struct. Geol. 4,
239e245.
fault nucleates. Carrillo, V.G., 2010. Eagle Ford Update/Barnett Lessons Learned: Developing
Unconventional Gas (DUG) Conference, Oct. 6, 2010 (San Antonio, TX).
Chapman, T.J., Williams, G.D., 1984. Displacement-distance methods in the analysis of
5. Conclusions fold-thrust structures and linked-fault systems, 141. Journal of the Geological Society,
London, pp. 121e128.
Analysis of small-scale thrust faults in interlayered limestone and Chapman, T.J., Williams, G.D., 1985. Strains developed in the hangingwalls of thrusts due
mudrock in the Ernst Member of the Boquillas Formation in Big Bend to their slip/propagation rate: a dislocation model: reply. J. Struct. Geol. 7, 759e762.
12 D.A. Ferrill et al. / Journal of Structural Geology 85 (2016) 1e12

Chapple, W.M., 1978. Mechanics of thin-skinned fold-and-thrust belts. Geol. Soc. Am. Morley, C.K., 1994. Fold-generated imbricates: examples from the Caledonides of Southern
Bull. 89, 1189e1198. Norway. J. Struct. Geol. 16, 619e631.
Chester, J.S., Chester, F.M., 1990. Fault-propagation folds above thrusts with constant dip. Moustafa, A.R., 1988. Structural Geology of Sierra del Carmen, Trans-Pecos Texas. 1: 48,000
J. Struct. Geol. 12, 903e910. scale geologic map: Geologic Quadrangle Map, 54, Bureau of Economic Geology. The
Dahlstrom, C.D.A., 1970. Structural geology in the eastern margin of the Canadian Rocky University of Texas at Austin, Austin, Texas.
Mountains. Bull. Can. Pet. Geol. 18, 332e406. Perry Jr., W.J., 1978. Sequential deformation in the central appalachians. Am. J. Sci. 278,
Donovan, A.D., Staerker, T.S., 2010. Sequence stratigraphy of the Eagle For d (Boquillas) 518e542.
Formation in the subsurface of south Texas an outcrops of west Texas. Gulf Coast Ramsay, J.G., Huber, M.I., 1983. The Techniques of Modern Structural Geology, vol. 1.
Assoc. Geol. Soc. Trans. 60, 861e899. Academic Press, p. 307.
Eisenstadt, G., DePaor, D.G., 1987. Alternative model of thrust-fault propagation. Geology Ramsay, J.G., 1992. Some geometric problems of ramp-flat thrust models. In: McClay, K.R.
15, 630e633. (Ed.), Thrust Tectonics. Chapman and Hall, pp. 191e200.
Erdlac Jr., R.H., 1994. Laramide paleostress trajectories from stylolites in the Big Bend Rich, J.L., 1934. Mechanics of low-angle overthrust faulting as illustrated by Cumberland thrust
region. In: Laroche, T.M., Viveiros, J.J. (Eds.), Structure and Tectonics of the Big block, Virginia, Kentucky, and Tennessee. AAPG Bull. 18, 1584e1596.
Bend and Southern Permian Basin. West Texas Geological Society 1994 Field Trip Rodgers, J., 1950. Mechanics of Appalachian folding as illustrated by Sequatchie anticline,
Guidebook, Publication 94-95, Texas, pp. 165e187. Tennessee and Alabama. AAPG Bull. 34, 672e681.
Ferrill, D.A., McGinnis, R.N., Morris, A.P., Smart, K.J., Sickmann, Z.T., Bentz, M., Snow, J.K., 1992. Large-magnitude Permian shortening and continental-margin tectonics in the
Lehrmann, D., Evans, M.A., 2014. Control of mechanical stratigraphy on bedrestricted southern Cordillera. Geol. Soc. Am. Bull. 104, 80e105.
jointing and normal faulting: Eagle Ford Formation, south-central Texas. U. S. A. Snow, J.K., Wernicke, B., 1989. Uniqueness of geological correlations: an example for the
AAPG Bull. 98, 2477e2506. Death Valley extended terrain. Geol. Soc. Am. Bull. 101, 1351e1362.
Ferrill, D.A., Morris, A.P., McGinnis, R.N., 2012. Extensional fault-propagation folding in Stewart, J.H., Anderson, R.E., Aranda-Gomez, J.J., Beard, L.S., Billingsley, G.H., Cather, S.M.,
mechanically layered rocks: the case against the frictional drag mechanism. Dilles, J.H., Dokka, R.K., Faulds, J.E., Ferrari, L., Grose, T.L.T., Henry, C.D., Janecke,
Tectonophysics 576e577, 78e85. S.U., Miller, D.M., Richard, S.M., Rowley, P.D., RoldanQuintana, J., Scott, R.B., Sears,
Ferrill, D.A., Morris, A.P., McGinnis, R.N., Smart, K.J., Ward, W.C., 2011. Fault zone J.W., Williams, V.S., 1998. Map showing Cenozoic tilt domains and associated structural
deformation and displacement partitioning in mechanically layered carbonates: the features, western North America e Plate 1. In: Faulds, J.E., Stewart, J.H. (Eds.),
Hidden Valley fault, central Texas. AAPG Bull. 95, 1383e1397. Accommodation Zones and Transfer Zones: The Regional Segmentation of the Basin and
Ferrill, D.A., Morris, A.P., Smart, K.J., 2007. Stratigraphic control on extensional fault Range Province. Geological Society of America Special Paper, p. 323.
propagation folding: Big Brushy Canyon monocline, Sierra del Carmen, Texas. In: Suppe, J., 1983. Geometry and kinematics of fault-bend folding. Am. J. Sci. 283, 684e721.
Jolley, S.J., Barr, D., Walsh, J.J., Knipe, R.J. (Eds.), Structurally Complex Reservoirs, Suppe, J., Medwedeff, D.A., 1990. Geometry and kinematics of fault-propagation folding.
292. Geological Society (London) Special Publication, pp. 203e217. Eclogae Geol. Helvetiae 83, 409e454.
Fischer, M.P., Woodward, N.B., Mitchell, M.M., 1992. The kinematics of break-thrust Tavani, S., Storti, F., Salvini, F., 2006. Double-edge fault-propagation folding: geometry and
folds. J. Struct. Geol. 14, 451e460. kinematics. J. Struct. Geol. 28, 19e35.
Flawn, P.T., Goldstein Jr., A., King, P.B., Weaver, C.E., 1961. The Ouachita System. Thorbjornsen, K.L., Dunne, W.M., 1997. Origin of a thrust-related fold: geometric vs kinematic
University of Texas at Austin. Bureau of Economic Geology Publication Number tests. J. Struct. Geol. 19, 303e319.
6120. Treadgold, G., McLain, B., Sinclair, S., Nicklin, D., 2010. Eagle Ford Shale prospecting with
Fleck, R.J., 1970. Tectonic style, magnitude, and age of deformation in the Sevier orogenic 3D seismic data within a tectonic and depositional system framework. Bull. South Tex.
belt in southern Nevada and eastern California. Geol. Soc. Am. Bull. 81, 1705e1720. Geol. Soc. LI 1, 19e28.
Gallup, W.B., 1951. Geology of Turner Valley oil and gas field, Alberta, Canada. AAPG Turner, K.J., Berry, M.E., Page, W.R., Lehman, T.M., Bohannon, R.G., Scott, R.B., Miggins,
Bull. 35, 797e821. D.P., Budahn, J.R., Cooper, R.W., Drenth, B.J., Anderson, E.D., Williams, V.S., 2011.
Hall, W.E., 1971. Geology of the panamint Butte Quadrangle, Inyo County, California and Geologic Map of Big Bend National Park. U. S. Geological Survey Scientific
Nevada. U. S. Geol. Surv. Bull. 1299, 67. Investigations Map 3142, scale 1, Texas, p. 84, 75,000, pamphlet.
Hancock, P.L., 1985. Brittle microtectonics: principles and practice. J. Struct. Geol. 7, Uzkeda, H., Poblet, J., Bulnes, M., 2010. A geometric and kinematic model for double-edge
437e457. propagating thrusts involving hangingwall and footwall folding. An example from the
Hardy, D., Allmendinger, R.W., 2011. Trishear: a review of kinematics, mechanics, and JacaePamplona Basin (Southern Pyrenees). Geol. J. 45, 506e520.
applications. In: McClay, K., Shaw, J.H., Suppe, J. (Eds.), Thrust Fault-related Wallace, W.K., Homza, T.X., 2004. Detachment folds versus fault-propagation folds, and their
Folding, 94. AAPG Memoir, pp. 95e119. truncation by thrust faults. In: McClay, K.R. (Ed.), Thrust Tectonics and Hydrocarbon
Hentz, T.F., Ruppel, S.C., 2010. Regional lithostratigraphy of the Eagle Ford Shale: Systems, 82. AAPG Memoir, pp. 324e355.
Maverick Basin to East Texas Basin. Gulf Coast Assoc. Geol. Soc. Trans. 60, 325e337. Welch, M.J., Knipe, R.J., Souque, C., Davies, R.K., 2009. A Quadshear kinematic model for
Hughes, A.N., Benesh, N.P., Shaw, J.H., 2014. Factors that control the development of folding and clay smear development in fault zones. Tectonophysics 471, 186e202.
fault-bend versus fault-propagation folds: insights from mechanical models based on Wernicke, B., Axen, G.J., Snow, J.K., 1988. Basin and Range extensional tectonics at the
the discrete element method (DEM). J. Struct. Geol. 68, 121e141. latitude of Las Vegas, Nevada. Geol. Soc. Am. Bull. 100, 1738e1757.
Hughes, A.N., Shaw, J.H., 2014. Fault displacement-distance relationships as in dicators of Wickham, J., 1995. Fault displacement-gradient folds and the structure at Lost Hills, California
contractional fault-related folding style. AAPG Bull. 98, 227e251. (U.S.A.). J. Struct. Geol. 17, 1293e1302.
Hughes, A.N., Shaw, J.H., 2015. Insights into the mechanics of fault-propagation folding Williams, G., Chapman, T., 1983. Strains developed in the hangingwalls of thrusts due to their
styles. Geol. Soc. Am. Bull. 127, 1752e1765. slip/propagation rate: a dislocation model. J. Struct. Geol. 5, 563e571.
Jamison, W.R., 1987. Geometric analysis of fold development in overthrust terranes. J.
Struct. Geol. 9, 207e219.
Katz, O., Reches, Z., Roegiers, J.C., 2000. Evaluation of mechanical rock properties using a
Schmidt Hammer. Int. J. Rock Mech. Min. Sci. 37, 723e728.
Kulander, B.R., Dean, S.L., 1986. Structure and tectonics of central and southern
Appalachian Valley and Ridge and Plateau Provinces, West Virginia and Vir ginia.
AAPG Bull. 70, 1674e1684.
Lehman, T.L., 1991. Sedimentation and tectonism in the Laramide Tornillo Basin of west
Texas. Sediment. Geol. 75, 9e28.
Link, T.E., 1949. Interpretations of foothills structures, Alberta, Canada. AAPG Bull. 33,
1475e1501.
Lock, B.E., Peschier, L., Whitcomb, N., 2010. The Eagle Ford (Boquillas Formation) of Val
Verde County, Texas e a window on the south Texas play. Gulf Coast Assoc. Geol. Soc.
Trans. 60, 419e434.
Maler, M.O., 1990. Dead Horse Graben: a west Texas accommodation zone. Tectonics 9,
1357e1368.
Maxwell, R.A., Lonsdale, J.T., Hazzard, R.T., Wilson, J.A., 1967. Geology of the Big Bend
National Park, Brewster County, 6711. University of Texas at Austin, Bureau of Economic
Geology Publication, Texas, p. 320.
McClay, K., 2011. Introduction to thrust fault-related folding. In: McClay, K., Shaw, J.H.,
Suppe, J. (Eds.), Thrust Fault-related Folding, 94. AAPG Memoir, pp. 1e19.
McConnell, D.A., Kattenhorn, S.A., Benner, L.M., 1997. Distribution of fault slip in outcrop-
scale fault-related folds, Appalachian Mountains. J. Struct. Geol. 19, 257e267.
McNaught, M.A., Mitra, G., 1993. A kinematic model for the origin of footwall syncline. J.
Struct. Geol. 15, 805e808.
Mitra, S., 1992. Balanced structural interpretation in fold and thrust bels. In: Mitra, S., Fisher,
G.W. (Eds.), Structural Geology of Fold and Thrust Belts. The Johns Hopkins University
Press, pp. 53e77.

You might also like