You are on page 1of 14

Journal of Structural Geology 58 (2014) 8e21

Contents lists available at ScienceDirect

Journal of Structural Geology


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

Evolution of the stress and strain fields in the Eastern Cordillera,


Colombia
Obi Egbue a, c, James Kellogg a, *, Hector Aguirre b, Carolina Torres a, c
a
Department of Earth and Ocean Sciences, University of South Carolina, Columbia, SC, USA
b
Equion Energy, Bogota, Colombia
c
BP, Houston, TX, USA

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

Article history: This work integrates stress data from Global Positioning System measurements and earthquake focal
Received 17 December 2012 mechanism solutions, with new borehole breakout and natural fracture system data to better understand
Received in revised form the complex interactions between the major tectonic plates in northwestern South America and to
28 August 2013
examine how the stress regime in the Eastern Cordillera and the Llanos foothills in Colombia has evolved
Accepted 17 October 2013
Available online 30 October 2013
through time. The dataset was used to generate an integrated stress map of the northern Andes and to
propose a model for stress evolution in the Eastern Cordillera. In the Cordillera, the primary present-day
maximum principal stress direction is WNWeESE to NWeSE, and is in the direction of maximum
Keywords:
Stress map
shortening in the mountain range. There is also a secondary maximum principal stress direction that is E
Global Positioning System eW to ENEeWSW, which is associated with the northeastward “escape” of the North Andean block,
Borehole breakouts relative to stable South America. In the Cupiagua hydrocarbon field, located in the Llanos foothills, the
Escape dominant NNEeSSW fractures are produced by the Panama arceNorth Andes collision and range-normal
Fractures compression. However, less well developed asymmetrical fractures oriented EeW to WSWeENE and
Eastern Cordillera NNWeSSE are also present, and may be related to pre-folding stresses in the foreland basin of the Central
Cordillera or to present-day shear associated with the northeastward “escape” of the north Andean block.
Our study results suggest that an important driver for orogenic deformation and changes in the stress
field at obliquely convergent subduction zone boundaries is the arrival of thickened crust, such as island
arcs and aseismic ridges, at the trench.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction work within the region, as well as the evolution of structural traps
and fracture systems for hydrocarbons. The present-day geology
Present-day stress in the northern Andes (Fig. 1) is a result of the shows the cumulative results of all the deformational events that
convergence of the South American, Caribbean, and Nazca plates have affected the region. To better understand both the neostress
and the Panama microplate. Global Positioning System (GPS) and paleostress regimes in the Eastern Cordillera, we have analyzed
measurements (Trenkamp et al., 2002) show that the North Andean data from a variety of stress indicators. The present-day stress
block is escaping toward the northeast relative to the stable South regime can be inferred from stress-induced borehole breakouts, as
American plate along a transpressive system of faults located along well as GPS vectors, and earthquake focal mechanism solutions. The
the Merida Andes, Eastern Cordillera, and Ecuadorian Andes paleostress is inferred from natural fracture systems and the
(Fig. 2). Simultaneously, in the Eastern Cordillera of Colombia, the orientation of structural features in the Llanos foothills. It can also
on-going Panama ArceSouth America collision is driving NWeSE be inferred from geologically dated measurements of north Andean
permanent shortening and rapid uplift of the range. Low-angle northeastward “escape”, the timing of the Panama arc collision, and
subduction of the buoyant Caribbean plate is also driving perma- age data for the uplift of the Eastern Cordillera. This study suggests
nent NWeSE shortening in the Merida Andes and Eastern Cordil- that the range-normal compressive stress field responsible for
lera. Knowledge of the evolving stress regime in the Eastern deforming rocks and producing the structural traps and fracture
Cordillera is key to understanding the geodynamic processes at systems in the Eastern Cordillera in the last 12 Ma has evolved to
include a range-parallel shear component in the last 2 Ma.
Paleostress analysis is an important tool in oil and gas explo-
* Corresponding author. ration, development, and reservoir simulation, constraining the
E-mail address: kellogg@sc.edu (J. Kellogg). development of subsurface fracture systems models. Subsurface

0191-8141/$ e see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jsg.2013.10.004
O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21 9

fracture systems in hydrocarbon reservoirs usually result from until later in the early 1980’s that they were attributed to the in-situ
paleostress that is unrelated to the present-day stress (Engelder, stress field (Gough and Bell, 1981; Hickman et al., 1982; Plumb,
1987). However, the neostress regime can strongly impact the 1982; Cox, 1983). A borehole breakout occurs when the stresses
permeability of fractures that formed in a different, older stress around the borehole exceed that required to cause compressive
regime. It is important therefore to understand how the stress failure of the borehole wall (Zoback et al., 1985). The development
regime has evolved through time. of intersecting conjugate shear planes causes spalling of the bore-
The North Andean margin provides an excellent Present-day hole wall, resulting in the enlargement of the well bore. In a vertical
example of an evolving stress field associated with oblique borehole, the wall shear stress is greatest in the direction of the
convergence and arc collision at a subduction zone. The strain minimum horizontal stress (Sh). The long axes of borehole break-
associated with oblique convergence at subduction zones is often outs therefore, are oriented approximately perpendicular to the
partitioned between trench-normal contraction on thrust faults maximum horizontal stress direction (SHmax) (Plumb and Hickman,
and trench-parallel strain on strike-slip faults (Fitch, 1972; Jarred, 1985).
1986; McCaffrey, 1994). In general, fore arc slivers form over the The borehole breakout data presented in this work (Fig. 1) is
region of interplate coupling and are driven along strike by the from four-arm caliper, high resolution dipmeter logs, and was
basal shear (Platt, 1993; McCaffrey et al., 2000). A volcanic arc can collected from 60 industry wells in the Eastern Cordillera and
help the partitioning process by localizing the margin-parallel surrounding areas. Each well was assigned a quality ranking based
shear strain in a zone of thermally induced lithospheric weakness on standard deviation, number of breakouts, and cumulative length
(Beck, 1983). In the North Andean margin, however, the Miocenee of broken out intervals using the rating scheme of Zoback and
Present collision with the Panama arc has increased coupling on the Zoback (1989). The breakout symbols plotted on the map repre-
plate boundary, inverted the back arc basin, and displaced the sent the location and direction of SHmax of each of the wells. The
margin-parallel shear strain to the back arc. In this study we bars show the direction of the mean azimuth of SHmax and the
combine GPS measurements and earthquake focal mechanisms length of the bars corresponds to the quality ranking of the
with geologic measurements of deformation and new borehole breakout data, which ranges from A to D. An “A” ranking represents
breakout and well fracture data to develop a model for the evolu- a well with many breakouts which have tightly clustered elonga-
tion of the strain field in the Eastern Cordillera over the last 14 Ma. tion azimuths. A “D” ranking represents a well with very few
Our study results suggest that the main driver for orogenic defor- breakouts, multiple azimuthal populations, and/or a large standard
mation and changes in the stress field was the arrival of thickened deviation for the mean azimuth. In general, borehole breakout
crust at the Ecuador and Colombia trenches, in particular, the orientations are bimodal, in the sense that data between 180 and
arrival of the Panama arc at 12e15 Ma and the Carnegie Ridge at 360 are equivalent to those between 0 and 180 . Borehole
2 Ma. breakouts in the region show significant scatter, but we propose
that the data fall into three major subsets by region: northern,
2. Present-day stress field data central, and southern. The northern subset is made up of wells
located above 6 N. Though a number of these wells show variable
2.1. Global Positioning System SHmax orientation, the majority have EeW trending SHmax. The
central subset is made up of wells located between 2 N and 6 N.
GPS studies are an important method for studying the kine- Most of the wells in this subset have WNWeESE trending SHmax.
matics of plate boundary zones, and intraplate deformation. The Wells in the southern segment (0 e2 N) have a NWeSE and less
velocity vectors used for this study (Fig. 1) are from the Central and commonly NNEeSSW trending SHmax. It is important to point out
South American (CASA) GPS project (Trenkamp et al., 2002) which that there are no clear patterns of azimuthal variations with depth
measured plate motions and crustal deformation of the four major in the individual wells or in the data subsets except for wells in the
plates in the region. The main results from the study can be sum- Llanos basin where the data shows a change in breakout orientation
marized as follows. In northern Colombia the Caribbean plate is at about 7800 ft (Torres Fleischmann, 1992). Below this depth, there
converging with stable South America in an EeSE direction at a rate are two main breakout population clusters. These correspond to
of 20  2 mm/a (Fig. 2). Panama is actively colliding with north- SHmax orientations of 110 and 070 .
western South America at a rate of approximately 25 mm/a. The
large eastward components of the GPS vectors in northern 2.3. Earthquake focal mechanism solutions
Colombia (Fig. 1, CART, MONT, RION, MZAL) suggest transfer of
motion from the Panama microplate. The oceanic Nazca plate and Some of the stress data in Fig. 1 are derived from single
the aseismic Carnegie Ridge are rapidly subducting at the Ecuadore earthquake focal mechanism solutions (FMS) from the World
Colombia trench at a rate of 58  2 mm/a relative to stable South Stress Map database (Reinecker et al., 2005). An earthquake focal
America. Approximately 50% of the NazcaeSouth America conver- mechanism solution is constructed using the radiation pattern of
gence is locked at the subduction interface, producing elastic strain seismic waves generated by an earthquake. The solution contains
in the overriding South American plate. GPS measurements also two perpendicular planes that represent the fault plane and the
show that the northern Andes is escaping northeastward at a rate of auxiliary plane, and divide the volume into compressional and
6  2 mm/a relative to stable South America (Fig. 1, e.g., BOGO, extensional quadrants. The orientation of the compressional (P),
PASTO, BUCM, MERI, VDUP, MZAL). intermediate (B), and extensional (T) axes are determined by the
orientations of these planes. It is important to point out, that the
2.2. Borehole breakouts moment tensor axes of earthquake focal mechanisms are not equal
to the principal stress axes. The only conclusion that can be made
Borehole breakouts indicate the horizontal stress field at inter- is that the maximum principal stress lies within the dilatational
mediate depths; usually less than 5 km. Borehole breakouts are quadrant of the focal mechanism (McKenzie, 1969). The principal
consistently oriented elongations in the wellbore cross-section that strain axes of earthquake focal mechanism solutions (P-axis, B-
form as a result of differential stress applied to an initially circular axis, and T-axis) are used as proxies for the orientation of the
borehole. Borehole breakouts were first observed by Cox (1970) in principal stresses axes s1, s2, and s3. To account for the un-
the Wapiabi shales of west-central Alberta. However, it was not certainties in this first-order approximation, the quality of the
10 O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21
O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21 11

Fig. 2. Schematic 3-D tectonic model for the northern Andes showing principal plate boundaries (after Egbue and Kellogg, 2010). Cross-section through 2 S latitude shows the
subducting Nazca plate and the detached North Andean block. Arrows show Present-day GPS-determined relative velocities of major plates (white) and the North Andes (black)
relative to stable South America (Trenkamp et al., 2002).

stress orientations derived from FMS in the World Stress Map 3. Paleostress field data
database is limited to a C-quality regardless of the size of the
earthquake and how well the focal mechanism is constrained. This The paleostress field in an orogenic belt can be inferred from the
quality ranking assumes a deviation of up to 25 . Focal mecha- geologic record, including fault and fold orientations, age-
nism solutions located in the vicinity of a plate boundary and with constrained fault displacements, rock deformation studies, and
kinematics similar to the plate boundary, are omitted and flagged natural fracture orientations. Here we summarize structural
as possible Plate Boundary Events (PBE) indicating that the P-, B-, models for the Eastern Cordillera and present new natural fracture
and T-axes of the focal mechanism solutions might predominantly data.
reflect the geometry and kinematics of the plate boundary, rather
than the orientation of the regional stress field. The accuracy of 3.1. Structure of the Eastern Cordillera
stress derivations from the FMS presented in this work are limited
by the fault-plane ambiguity and the coefficient of friction (Barth The Eastern Cordillera has been generally interpreted as a wide
et al., 2008). Mesozoic extensional basin that evolved into a foreland basin for
SHmax orientations determined from FMS shown in Fig. 1 appear the Central Cordillera, and was subsequently inverted by Andean
to fall into two main subsets: showing NWeSE and ENEeWSW compression and orogeny (Colletta et al., 1990; Dengo and Covey,
maximum compressive stress directions (Ego et al., 1996; Corredor, 1993; Cooper et al., 1995). Based on a retrodeformed cross-
2003; Cortés and Angelier, 2005). The NWeSE compression has section through the Eastern Cordillera, Colletta et al. (1990) pro-
been associated with CaribbeaneSouth American convergence and posed that the Cordillera was formed by the inversion of two deep
PanamaeSouth America collision (Kellogg and Vega, 1995; Taboada Upper JurassiceLower Cretaceous basins during Mio-Pliocene
et al., 2000; Trenkamp et al., 2002; Corredor, 2003; Colmenares times. In this model, the structure is controlled by low angle
and Zoback, 2003; Acosta et al., 2004), and the ENEeWSW thrusts and reactivated normal faults that constitute frontal ramps.
compression has been ascribed to NazcaeSouth America conver- They estimated a total shortening of at least 105 km with a
gence and Carnegie Ridge subduction at the ColombiaeEcuador decollement at a depth of about 20 km. Toro et al. (2004) obtained
trench (Ego et al., 1996; Gutscher et al., 1999; Trenkamp et al., 2002; similar results along two partial volume-balanced transects of the
Corredor, 2003; Colmenares and Zoback, 2003; Egbue and Kellogg, Cordillera. Dengo and Covey (1993) proposed that basement-
2010). detached (“thin-skinned”) shortening was followed by uplift on

Fig. 1. Stress map of the northern Andes, showing GPS vectors (Trenkamp et al., 2002), and inferred directions of maximum horizontal stress from borehole breakouts (This study
and Torres Fleischmann, 1992) and earthquake focal mechanisms (Reinecker et al., 2005). GPS velocity vectors are relative to stable South America at 95% confidence level. The
lengths of the stress bars indicate data quality.
12 O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21

high-angle basement involved reverse faults (“thick-skinned”) (2012) modeled approximately 4 km (9%) NWeSE shortening in
deformation and estimated approximately 150 km of shortening the Late Oligocene (25 Ma) followed by 13e19 km (30e40%) NWe
(40%) from a regional retrodeformable cross-section. Cooper et al. SE shortening in the last 10 million years.
(1995) proposed that deformation in the Llanos foothills was pre- Based on paleostress indicators (fault and fold orientations),
dominantly inversion of pre-existing extensional faults by Cortés et al. (2005) present evidence of ENEeWSW directed
basement-involved “thick skinned” listric reverse faults. They compression and shear stresses in the present day western flank
estimated only 68 km of shortening. and center of the Eastern Cordillera during the MaastrichtianeLate
Egbue and Kellogg (2012) studied the three-dimensional Paleocene, prior to the uplift of the Eastern Cordillera. Geologic
structural evolution of the Piedemonte block in the NEeSW indicators of ENEeWSW compression and right-lateral shear in the
trending central Llanos foothills, the frontal thrust zone along the Eastern Cordillera in the last 2 Ma are largely obscured by the
eastern flank of the Eastern Cordillera (Fig. 3). NWeSE directed ongoing NWeSE compression. The timing of this northeastward
compression has formed a duplex zone containing as many as five “escape” is inferred from geological observations in Ecuador,
east-verging stacked thrust sheets. Using 2-D and 3-D seismic southern Colombia, and Venezuela (compiled in Egbue and Kellogg,
reflection data, surface geology, and well data, Egbue and Kellogg 2010). The ongoing ENEeWSW compression is also evidenced by

Fig. 3. Geologic map of the Llanos foothills on the eastern flank of the Eastern Cordillera showing locations of several major hydrocarbon features in the Llanos foothills, including
the Piedemonte block and the Cupiagua and Cusiana fields (BP Exploration Company Colombia). Insets show the locations of Figs. 5 and 11.
O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21 13

present-day GPS measurements, earthquake focal mechanisms, natural fractures are present at different stratigraphic intervals
and bore-hole breakout data. including the main reservoirs. Fractures are a modifier of reservoirs,
enhancing permeability when they occur open and connected, and
3.2. Natural fractures reducing permeability when they are closed. In the reservoir in-
tervals, fractures with a low acoustic amplitude and high transit
Natural fractures are typically present in all rocks to varying time (i.e. dark on UBI images) are dominant, and are interpreted as
degrees. In fold and thrust belt regions, the rocks are even more being open or partially open at the borehole wall (Adams et al.,
intensely fractured. These fractures are developed and/or modified 1999). A high transit time suggests an aperture at the borehole
during deformation in well consolidated rocks. Most of the tight, wall and a mud or clay fill which gives rise to low acoustic
well-cemented reservoirs, like those in the flanks of the Eastern impedance. However, clay smeared fractures are rare in cores from
Cordillera, have natural fractures which are important features that Cupiagua; hence the fractures are interpreted as being filled with
control fluid-reservoir dynamics. Subsequent stress regimes can drilling mud. Fractures with high acoustic amplitude are rare. UBI
strongly affect the permeability of ancient fractures that formed in images show acoustic amplitude contrast which is a response to
a different, older stress regime. The propagation paths for fractures changes in acoustic impedance at the borehole wall, which in
in a geologic setting are strongly influenced by the state of stress clastic sequences is related mainly to changes in matrix porosity
and are only weakly dependent on the rock fabric (Olson and and lithology. Within the Cupiagua field, open fractures commonly
Pollard, 1989). Vertical fractures propagate normal to the least occur in proximity to cataclastic faults suggesting that fracture
principal stress, thus, they follow the path of the stress field present propagation has a genetic relationship with major deformation
at the time of their propagation (Engelder and Geiser, 1980). By episodes. The mean fracture density varies within the reservoir
examining a map of fracture-set orientation, we can infer the dif- units, and ranges between 0.11e0.62 fractures per foot (Adams
ferential stress that was active during their propagation. et al., 1999). A stratigraphic summary diagram for the Llanos foot-
In the Cupiagua hydrocarbon field (Fig. 3) on the eastern flank of hills (Fig. 4) shows the key lithostratigraphic units/formations,
the Eastern Cordillera, cores and image logs (Ultrasonic Borehole reservoir units, unconformities, and detachment levels. In general,
Imager-UBI and Fullbore Formation MicroImager-FMI) show that the Barco formation is more intensely fractured than the Mirador

Fig. 4. Stratigraphic summary diagram for the Llanos foothills showing typical gamma ray log response, the key lithostratigraphic units/formations, proven reservoir rocks, source
rocks, and regional seals. On the left is a schematic cross section for the structure of the Cupiagua field. Stratigraphic information is from BP Exploration Company Colombia.
14 O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21

and Guadalupe Formations. It is important to point out that the angles below 15,000 ft. Stearns (1967) observed that infold struc-
variations in fracture density may not be solely a result of geological tures in the western United States, conjugate and tensile fractures
factors. It is also dependent on variations in UBI image quality. form sets that are oriented in varying directions relative to bedding
Based on log images (UBI and FMI) from wells in the Cupiagua and the fold hinge line (Fig. 7). In Cupiagua, a NWeSE trending
field, we recognize three main families of fractures in the Cupiagua fracture set is widespread throughout the field, and is dominant in
field (Fig. 5). The data was classified, correlated, and plotted as rose the southern portion of the field (Fig. 5). These fractures are aligned
diagrams to group fracture events into clusters/families. A fracture parallel to the present-day regional stress maximum. An EeW to
dip vs depth plot from UBI for the wells (Fig. 6) shows that most of WNWeESE family of fractures is approximately orthogonal to the
the natural fractures present in the reservoir between 12,000 and axis of the structure. This family of fractures is dominant in the
16,000 ft are steeply dipping to subvertical (40e90 ) suggesting northern segment of the field, but these fractures also occur in the
that the maximum principal stress directions are subhorizontal. central and southern segments. This fracture set corresponds to
Only one anomalous well shows a wide dispersion of fracture dip Stearns’ fracture type 1 (Fig. 7a). A third NNEeSSW family of

Fig. 5. Fracture strike rose diagram plot for the northern, central, and southern Cupiagua field. The Cupiagua hydrocarbon field is located in Fig. 3. Black dots show well locations.
O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21 15

Fig. 6. Fracture dip/depth plot from UBI for Cupiagua field wells shown in Fig. 5. Different symbols represent different wells. Note that most of the natural fractures present in the
reservoir between 12,000 and 16,000 ft depths are steeply dipping to subvertical (40e90 ).

fractures is common in the central portion of the field (Fig. 5), and is
also present in the southern segment. These type 2 fractures run
parallel to the axis of the Cupiagua structure and occur in both the
crest and flank regions.

4. Discussion

4.1. Present-day stress

The regional stress map (Fig. 1) shows that the primary


present-day maximum principal stress direction in the Eastern
Cordillera is WNWeESE to NWeSE. This direction is roughly
orthogonal to the mountain range and in the direction of
maximum shortening. Colmenares and Zoback (2003) proposed
that the negatively buoyant Caribbean slab may be driving the
rotation of SHmax to the ESE in the direction of relative subduction
under the northern Andes. The WNWeESE stress direction is also
the PanamaeNorth Andes relative convergence vector when cor-
rected for the 6 mm/a northeastward “escape” of the North Andes,
and the magnitude of the shortening required to build the Eastern
Cordillera (68e150 km, 3.1 Structure of the Eastern Cordillera),
suggests that most of the continuing shortening and uplift of the
range is associated with the ongoing PanamaeNorth Andes colli-
sion. Close to the Panama block indenter on the western side of
the northern Andes, the collision-related stresses are particularly
strong. For example, south of MZAL, the 1999 Armenia earthquake
(Fig. 1, focal mechanism at 4.4 N, 284.3 E) may have been
localized left-lateral slip on the NNE-trending Romeral fault sys-
tem related to the PanamaeNorth Andes collision (Pulido, 2003).
We interpret this as Panama collision-produced shear on the
broad western flank of the northeastward “escaping” North Andes
block.
A secondary present-day maximum principal stress direction in
the Eastern Cordillera is EeW to ENEeWSW. This is particularly
pronounced in the northern termination of the Eastern Cordillera,
the Bocono fault system in Venezuela, and in southern Colombia
and Ecuador. We propose that this stress direction is associated
with the subduction of the Carnegie Ridge and the northeastward
Fig. 7. Schematic diagrams showing (a) fractures types associated with folding, after “escape” of the North Andes relative to stable South America. The
Stearns (1967) and (b) an example of an asymmetrical fracture unrelated to folding. stress is particularly noticeable at the northern end of the Eastern
16 O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21

Fig. 8. Eastern Cordillera uplift, shortening, and escape history (a) Uplift history of the Eastern Cordillera. Estimated paleoelevation of the Eastern Cordillera (black squares with
error bars after Wijninga, 1996; Gregory-Wodzicki, 2000) and tectonic uplift (red lines). Tectonic uplift is estimated from present day total uplift (10 km) for erosion rates of 0.5 and
1.0 mm/a. b) Estimated shortening rates for the Eastern Cordillera since PanamaeSouth America collision began and northeastward “escape” rates for the North Andean block
(Egbue and Kellogg, 2010). Shortening rates proportional to tectonic uplift rates are estimated for 60 km and 120 km shortening. The present-day PanamaeSouth America
convergence rate is approximately 30 mm/yr (Trenkamp et al., 2002). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)

Cordillera where a left-stepping restraining bend produces per- 4.2.1. Oligocene stresses
manent compressional deformation. By Oligocene time, uplift and shortening of the Central
Cordillera led to uplift in the foreland basin (now the western
foothills of the Eastern Cordillera) (Horton et al., 2010; Moreno
4.2. Paleostress et al., 2011). Apatite fission-track and zircon (UeTh)/He 30 (ZHe)
ages support an Oligocene age for the onset of deformation in the
During the Late Jurassic and Cretaceous periods, the area of the central axis of the Eastern Cordillera (Mora et al., 2010a; Saylor
present-day Eastern Cordillera was a rift basin with approximately et al., 2012). Egbue and Kellogg (2012) estimate Late Oligocene
WNWeESE oriented extension (minimum principal stress) (Ojeda, shortening as 20% of the total Cenozoic shortening in the eastern
1996; Sarmiento-Rojas et al., 2006). Based on low-temperature foothills of the Eastern Cordillera. Hossack (1999) reported the
thermochronometry and seismic reflection data from the middle presence of asymmetrical fracture sets (Fig. 7b) that appear to
Magdalena Valley and field paleostress indicators, Parra et al. predate folding in the Cupiagua structure. This was done by
(2012) and Cortés et al. (2005) claim evidence for Paleocene to removing those fractures that are related to folding from the
Eocene ENEeWSW contraction on the western flank of the present fractures dataset. Prior to the subtraction, the fracture data was
Eastern Cordillera. rotated so that the bedding is unfolded to the horizontal. Type 1
O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21 17

Fig. 9. Schematic block diagram showing (a) early compressive fracture generation and (b) late stage shear fractures superimposed on earlier fracture system.

and type 2 fracture sets were removed from all the fracture data, a 4.2.2. Andean stresses (Miocene to Present)
process which is similar to backstripping. Two less well developed Paleobotanical data (Fig. 8a; Wijninga, 1996; Gregory-Wodzicki,
fracture sets remain that are orthogonal: a WSWeENE and a 2000) and fission-track ages (Mora, 2007; Mora et al., 2010b)
NNWeSSE trend. Since these fractures remain after the bedding indicate that most of the uplift in the central and eastern flank of
dips on the different parts of the fold are unfolded back to hori- the Eastern Cordillera occurred in the last 10 Ma. Tectonic uplift
zontal, they are interpreted as predating folding of the Cupiagua (Fig. 8a) is estimated from present-day total uplift and paleo-
structure. Orthogonal fracture sets of this type are not uncommon, elevation data. Elevation is shown in black with present mean
and are typically found in undeformed foreland areas in front of elevation just over 3 km and estimated paleoelevation based on
thrust belts. Some examples are the fracture sets of the Appala- paleobotanical and geomorphological data (black squares with er-
chian Plateau (Engelder and Geiser, 1980), the Alberta foreland ror bars after Wijninga, 1996; Gregory-Wodzicki, 2000). Tectonic
(Babcock, 1974), and the European Platform from the Alps to as far uplift (red line in online version) is estimated using present day
as southern England (Bevan and Hancock, 1986). In all cases, a total structural relief (10 km) and assuming erosion rates of 0.5 and
main fracture set is usually orthogonal to the thrust belt and a 1.0 mm/a. Neogene exhumation history from apatite fission track
second set lies at right angles to the main set. Younes and Engelder data and low-T thermochronology (Mora, 2007; Parra et al., 2009)
(1999) have shown that the earliest fractures in the Appalachian supports the continued rapid tectonic uplift proposed for the last
plateau developed prior to a major phase of layer parallel short- 3 Ma. Using apatite fission-track data, Mora et al. (2010a,b) esti-
ening. Continued shortening and detachment sliding of the cover mated exhumation rates of 1e1.5 mm/a for compressional struc-
rocks developed a new set of fractures that are slightly rotated tures in the eastern foothills of the Eastern Cordillera over the last
compared to the earlier set. If we assume that the orthogonal 3 Ma. Wittmann et al. (2011) used cosmogenic nuclide-based
fractures in the Cupiagua field are analogous to the Appalachian measurements to estimate denudation rates of 0.49e1.2 mm/a for
fracture sets, then it is possible that they could have formed the Ecuadorian Andes. Note that the uplift curve for an erosion rate
during an earlier phase of layer-parallel shortening (Hossack, of 0.5 mm/a shows tectonic uplift in the Eastern Cordillera
1999). On the other hand, this EeW to ENEeWSW stress orien- distributed over at least 10 Ma.
tation is the same as that observed in the present-day stress field The shortening rates for the Eastern Cordillera (Fig. 8b; Egbue
associated with the northeastward “escape” of the northern and Kellogg, 2012) since 12 Ma (the initiation of the Panamae
Andes, and the two may therefore be confused. South America collision) are estimated from the tectonic uplift rates
18 O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21

Fig. 10. Model for fracture development in the Cupiagua field (after Hossack, 1999) (a) fracture sets orthogonal and perpendicular to the early thrust belt. (b) ENEeWSW superposed
fracture set related to propagating thrust front or present-day shear.

(Fig. 8a) and constrained to fit the present GPS measured rate, Present) of deformation (Fig. 9a), thrusting and folding in the re-
12  4 mm/a (Trenkamp et al., 2002). Shortening rates (v) are gion generated most of the Eastern Cordillera and foothills struc-
assumed to be proportional to tectonic uplift rates (u): v ¼ ku, tures orthogonal to the main stress and developing fold-related
where k is a dimensionless proportionality constant. Shortening conjugate fractures. Two sets of fractures, symmetrically orientated
rates are shown for a total shortening of 120 km in the last 12 Ma to the fold hinges of the Cupiagua structure, form a strike-parallel
(0.8  150 km; Dengo and Covey, 1993; Egbue and Kellogg, 2012). type 2 fracture set and a strike-normal type 1 fracture set. A rota-
For 120 km shortening, k ¼ 18. Shortening rates are also shown for tion of the shortening direction to a direction normal to the Cusiana
60 km shortening, the minimum published estimate for the Eastern (and Cupiagua structures (Fig. 3)) probably occurred as the thrust
Cordillera (0.8  68 km; Cooper et al., 1995), where k ¼ 9. North- front advanced into the Llanos foothills (Hossack, 1999) (Fig. 10).
eastward escape rates for the North Andean block (Fig. 8b) are from Subsequent thrusting and folding generated the type 1 and type 2
the compilation by Egbue and Kellogg (2010). The present-day fractures which were then superposed onto the earlier orthogonal
PanamaeSouth America convergence rate is approximately set. The orientation of structures in the Piedemonte hydrocarbon
30 mm/yr (Trenkamp et al., 2002). Note that range-parallel shear block located northeast of the Cupiagua field, as seen on 3-D
“escape” has become an important factor in the last 2 Ma, but seismic reflection depth slices (Fig. 11), shows evidence of this
range-normal shortening rates exceed the shear rates, and total late stage thrusting. The structures are oriented perpendicular to
shortening is 3e6 times as great as the shear displacement in the the NWeSE present-day maximum principal stress direction in the
last 12 Ma. Eastern Cordillera.
About 12e15 Ma the Panama arc began colliding with the
northern Andes (Coates et al., 2004; Montes et al., 2012), and the 4.2.3. Early Pleistocene range-parallel shear
Eastern Cordillera became a major sediment source (Gregory- By Early Pleistocene (Figs. 8 and 9b), in addition to the
Wodzicki, 2000), although some sediments were being supplied compressional stress regime, a NEeSW strike-slip component was
to the Eastern Foothills area as early as the Oligocene (Mora et al., introduced as the northern Andes began to “escape”. About 2 Ma,
2010a; Saylor et al., 2012). During the Andean stage (Miocene to the aseismic Carnegie ridge, which was formed by the Galapagos
O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21 19

Fig. 11. 3-D seismic reflection depth slices through the Piedemonte block showing the orientation of major structural features at (a) 2270 m, (b) 4300 m, and (c) 5280 m.
20 O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21

hotspot, arrived at the ColombiaeEcuador trench (Lonsdale and southern coast of Cabo San Lorenzo (Manabí, Ecuador). J. South Am. Earth
Sci. 16, 633e648.
Klitgord, 1978; Pedoja, 2003; Cantalamessa and Di Celma, 2004),
Coates, A.G., Collins, L.S., Aubry, M.-P., Berggren, W.A., 2004. The geology of the
and initiated the northeastward “escape” of the northern Andes Darien, Panama, and the late Miocene-Pliocene collision of the Panama arc with
(Egbue and Kellogg, 2010). This resulted in the development of northwestern South America. Geol. Soc. Am. Bull. 116, 1327e1344. http://
asymmetrical shear fractures oriented EeW to WSWeENE and dx.doi.org/10.1130/B25275.1.
Colletta, B., Hebrard, F., Letouzey, J., Werner, P., Rudkiewickz, J., 1990. Tectonic style
NNWeSSE similar to the fractures interpreted by Hossack (1999) as and crustal structure of the Eastern Cordillera (Colombia) from a balanced cross
predating the folding. Presently, in the Eastern Cordillera the section. In: Letouzey, J. (Ed.), Petroleum and Tectonics in Mobile Belts. Editions
escape rate is almost as great as the rate of shortening due to the Technip, Paris, pp. 81e100.
Colmenares, L., Zoback, M.D., 2003. Stress field and seismotectonics of northern
collision with Panama (Trenkamp et al., 2002; Egbue and Kellogg, South America. Geology 31 (8), 721e724.
2012). However, since compressive strain results in higher Cooper, M.A., Addison, F.T., Álvarez, R., Coral, M., Graham, R.H., Hayward, S.H.,
stresses than shear strain, range normal compression still domi- Martinez, J., Naar, J., Peñas, R., Pulham, A.J., Taborda, A., 1995. Basin develop-
ment and tectonic history of the Llanos Basin, Eastern Cordillera, and Middle
nates the stress field in the Eastern Cordillera. Magdalena Valley: Colombia. Am. Assoc. Pet. Geol. Bull. 79 (10), 1421e1443.
Corredor, F., 2003. Seismic strain rates and distributed continental deformation in
the northern Andes and three-dimensional seismotectonics of northwestern
5. Conclusions
South America. Tectonophysics 372 (3e4), 147e166.
Cortés, M., Angelier, J., 2005. Current states of stress in the northern Andes as
Borehole breakouts and seismic focal mechanisms indicate two indicated by focal mechanisms of earthquakes. Tectonophysics 403, 29e58.
Cortés, M., Angelier, J., Colleta, B., 2005. Paleostress evolution of the northern Andes
present-day principal stress directions in the Eastern Cordillera;
(Eastern Cordillera of Colombia): implications on plate kinematics of the South
WNWeESE to NWeSE, aligned with the GPS measured Panamae Caribbean region. Tectonics 24, TC1008. http://dx.doi.org/10.1029/
North Andes collision, and EeW to WSWeENE, aligned with the 2003TC001551.
northeastward “escape” of the North Andes. The Panama collision Cox, J.W., 1970. The high resolution dipmeter reveals dip-related borehole and
formation characteristics. In: Transactions, 11th Annual Logging Symposium,
has produced most of the folding, shortening, and uplift of the Los Angeles, pp. D1eD25.
Eastern Cordillera. The dominant fracture systems observed in the Cox, J.W., 1983. Long axis orientation in elongated boreholes and correlation with
Cupiagua fold structures on the eastern flank of the Eastern rock stress data. In: 24th Annual Logging Symposium, Calgary, pp. 1e17.
Dengo, C., Covey, M., 1993. Structure of the Eastern Cordillera of Colombia: impli-
Cordillera are produced by the collision and range-normal cations for trap styles and regional tectonics. Am. Assoc. Pet. Geol. Bull. 77,
compression. Less well developed asymmetrical shear fractures 1315e1337.
oriented EeW to WSWeENE and NNWeSSE may be related to pre- Egbue, O., Kellogg, J., 2010. Pleistocene to present North Andean “escape”. Tecto-
nophysics 489, 248e257 http://dx.doi.org/10.1016/j.tecto.2010.04.021.
folding stresses in the foreland basin of the Central Cordillera Egbue, O., Kellogg, J., 2012. Three-dimensional structural evolution and kinematics
(future Eastern Cordillera) or to present-day shear associated with of the Piedemonte Llanero, Central Llanos foothills, Eastern Cordillera,
the northeastward “escape” of the north Andean block. However, Colombia. J. South Am. Earth Sci.. http://dx.doi.org/10.1016/
j.jsames.2012.04.012.
range-normal compression still dominates the stress field in the
Ego, F., Sebriera, M., Lavenuc, A., Yepes, H., Egues, A., 1996. Quaternary state of stress
Eastern Cordillera. Our study suggests that the critical driver for in the northern Andes and the restraining bend model for the Ecuadorian
orogenic deformation at transpressive plate boundaries is the Andes. Tectonophysics 259 (1e3), 101e116.
Engelder, T., 1987. Joints and shear fractures in rock. In: Atkinson, B.K. (Ed.), Fracture
increased coupling associated with the subduction of aseismic
Mechanics of Rock. Academic Press, London, pp. 27e69.
ridges and island arc collisions. Both range-normal shortening and Engelder, T., Geiser, P., 1980. On the use of regional joint sets as trajectories of
range-parallel shear displacement must be considered when esti- paleostress fields during the development of the Appalachian plateau, New
mating the earthquake hazard for the Eastern Cordillera and the York. J. Geophys. Res. 85, 6319e6341.
Fitch, T.J., 1972. Plate convergence, transcurrent faults, and internal deformation
populous Bogota area. adjacent to southeast Asia and the western Pacific. J. Geophys. Res. 84, 4432e
4460.
Gough, D.I., Bell, J.S., 1981. Stress orientations from oil-well fractures in Alberta and
Acknowledgments Texas. Can. J. Earth Sci. 18, 638e645.
Gregory-Wodzicki, K.M., 2000. Andean paleoelevation estimates: a review and
The authors would like to thank Editor Cees Passchier, Joel critique. Geol. Soc. Am. Bull. 112, 1091e1105.
Gutscher, M.-A., Malavieille, J., Lallemand, S., Collot, J.-Y., 1999. Tectonic segmen-
Saylor, and an anonymous reviewer, for very constructive reviews.
tation of the North Andean margin: impact of the Carnegie Ridge collision.
We thank BP Colombia and Equion Energy Colombia for making the Earth Planet. Sci. Lett. 168, 255e270.
fracture data from Cupiagua field and the 3-D seismic reflection Hickman, S.H., Healy, J.J., Zoback, M.D., Svitek, J.F., Bretcher, J.E., 1982. In situ stress,
borehole elongation, and natural fracture distribution, at depth in central New
data available, as well as providing partial support for the first
York State [abs.]. Eos (Trans., Am. Geophys. Union) 63, 1118.
author’s research at the University of South Carolina. The CASA Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., Stockli, D.F.,
project was supported by NSF Grants EAR-8617485, EAR-8904657, 2010. Linking sedimentation in the northern Andes to basement configuration,
and NASA. The borehole breakout data was supplied by Ecopetrol. Mesozoic extension, and Cenozoic shortening: evidence from detrital zircon U-
Pb ages, Eastern Cordillera, Colombia. Geol. Soc. Am. Bull. 122 (9/10), 1423e
Figs. 1 and 3 were created with Generic Mapping Tool (GMT) 1442,. http://dx.doi.org/10.1130/B30118.1.
(Wessel and Smith, 1998). Hossack, J., 1999. An Analysis of the Geometry, Orientation, and Structural Devel-
opment of the Schlumberger UBI Fracture Sets in Cupiagua and Cusiana Wells.
Internal Document. BP Exploration, Colombia, p. 33.
References Jarred, R.D., 1986. Relations among subduction parameters. Rev. Geophys. 24,
217e284.
Acosta, J., Lonergan, L., Coward, M.P., 2004. Oblique transpression in the western Kellogg, J.N., Vega, V., 1995. Tectonic development of Panama, Costa Rica, and the
thrust front of the Eastern Cordillera. J. South Am. Earth Sci. 17, 181e194. http:// Colombian Andes: constraints from Global Positioning System geodetic studies
dx.doi.org/10.1016/j.jsames.2004.06.002. and gravity. In: Mann, P. (Ed.), Geologic and Tectonic Development of the
Adams, J., Styles, A., Garcia-Carballido, C., 1999. Fracture Characterization from UBI Caribbean Plate Boundary in southern Central America, Geological Society of
Images, Cupiagua Field: Pilot Study. Internal Document. BP Exploration, America Special Paper 295, pp. 75e90.
Colombia, p. 125. Lonsdale, P., Klitgord, K.D., 1978. Structure and tectonic history of the eastern
Babcock, E., 1974. Jointing in central Alberta. Can. J. Earth Sci. 11, 1181e1186. Panama Basin. Geol. Soc. Am. Bull. 89, 981e999.
Barth, A., Reinecker, J., Heidbach, O., 2008. Stress Derivation from Earthquake Focal McCaffrey, R., 1994. Global variability in subduction thrust zone-forearc systems.
Mechanisms. World Stress Map Project e Guidelines, p. 12. Pure Appl. Geophys. 141, 173e224.
Beck, M.E., 1983. On the mechanism of tectonic transport in zones of oblique McCaffrey, R., Zwick, P.C., Bock, Y., Prawirodirdjo, L., Genrich, J.F., Stevens, C.W.,
subduction. Tectonophysics 93, 1e11. Puntodewo, S.S.O., Subarya, C., 2000. Strain partitioning during oblique plate
Bevan, T.G., Hancock, P.L., 1986. A late Cenozoic regional microfracture system in convergence in northern Sumatra: geodetic and seismologic constraints and
southern England and northern France. J. Geol. Soc. Lond. 143, 355e362. numerical modeling. J. Geophys. Res. 105 (B12), 28,363e28,376.
Cantalamessa, G., Di Celma, C., 2004. Origin and chronology of Pleistocene ma- McKenzie, D.P., 1969. The relation between fault plane solutions for earthquakes
rine terraces of Isla dela Plata and of flat, gently dipping surfaces of the and the directions of the principal stress. Bull. Seismol. Soc. Am. 59, 591e601.
O. Egbue et al. / Journal of Structural Geology 58 (2014) 8e21 21

Montes, C., Bayona, G., Cardona, A., Buchs, D.M., Silva, C.A., Morón, S., Hoyos, N., Reinecker, J., Heidbach, O., Tingay, M., Sperner, B., Müller, B., 2005. The 2005 Release
Ramírez, D.A., Jaramillo, C.A., Valencia, V., 2012. Arc-continent collision and of the World Stress Map online at: www.world-stress-map.org.
orocline formation: closing of the Central American seaway. J. Geophys. Res. 117, Sarmiento-Rojas, L.F., Van Wess, J.D., Cloetingh, S., 2006. Mesozoic transtensional
B04105. http://dx.doi.org/10.1029/2011JB008959. basin history of the Eastern Cordillera, Colombian Andes: inferences from
Mora, A., 2007. Inversion Tectonics and Exhumation Processes in the Eastern tectonic models. J. South Am. Earth Sci. 21 (4), 383e411. http://dx.doi.org/
Cordillera of Colombia (Ph.D. thesis). University of Potsdam, Potsdam, 133 p. 10.1016/j.jsames.2006.07.003.
Mora, A., Horton, B.K., Mesa, A., Rubiano, J., Ketcham, R.A., Parra, M., Blanco, V., Saylor, J., Horton, B., Stockli, D., Mora, A., Corredor, J., 2012. Structural and ther-
Garcia, D., Stockli, D.F., 2010a. Migration of Cenozoic deformation in the Eastern mochronological evidence for Paleogene basement-involved shortening in the
Cordillera of Colombia interpreted from fission track results and structural re- axial Eastern Cordillera, Colombia. J. South Am. Earth Sci. 39, 202e215.
lationships: implications for petroleum systems. AAPG Bull. 94 (10), 1543e1580. Stearns, D.W., 1967. Certain aspects of fracture in naturally deformed rocks. In:
http://dx.doi.org/10.1306/01051009111. Riecker, R.E. (Ed.), National Science Foundation Advanced Science Seminar in
Mora, A., Parra, M., Strecker, M.R., Sobel, E.R., Zeilinger, G., Jaramillo, C., Da Silva, S., Rock Mechanics: Special Report: Air Force Cambridge Research Lab, Bedford,
Blanco, M., 2010b. The eastern foothills of the Eastern Cordillera of Colombia: an Massachusetts, pp. 97e118.
example of multiple factors controlling structural styles and active tectonics. Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Hervé, P., Harmen, B., Olaya, J.,
Geol. Soc. Am. Bull. 112, 1846e1864. http://dx.doi.org/10.1130/B30033.1. Rivera, C., 2000. Geodynamics of the northern Andes: subductions and intra-
Moreno, C., Horton, B., Caballero, V., Mora, A., Parra, M., Sierra, J., 2011. Depositional continental deformation (Colombia). Tectonics 19 (5), 787e813.
and provenance record of the Paleogene transition from foreland to hinterland Toro, J., Roure, F., Bordas-Le Floch, N., Le Cornec-Lance, S., Sassi, W., 2004. Thermal
basin evolution during Andean orogenesis, northern Middle Magdalena Valley and kinematic evolution of the Eastern Cordillera fold and thrust belt,
Basin, Colombia. J. South Am. Earth Sci. 32 (3), 246e263. Colombia. In: Swennen, R., Roure, F., Granath, J.W. (Eds.), Deformation, Fluid
Ojeda, G., 1996. Present Day Retrodeformed Structure of the Eastern Cordillera Flow, and Reservoir Appraisal in Foreland Fold and Thrust Belts, AAPG Hedberg
Cretaceous Basin, Colombia (M.S. thesis). University of South Carolina, Series, no. 1, pp. 79e115.
Columbia, 57 p. Torres Fleischmann, C., 1992. In-situ Stress in the North Andes from Borehole
Olson, J., Pollard, D.D., 1989. Inferring paleostresses from natural fracture patterns: a Breakout Analysis (MS. thesis). University of South Carolina, Columbia, 153 p.
new method. Geology 17, 345e348. Trenkamp, R., Kellogg, J.N., Freymuller, J.T., Mora, H.P., 2002. Wide plate margin
Parra, M., Mora, A., Sobel, E.R., Strecker, M.R., González, R., 2009. Episodic orogenic deformation, southern Central America and northwestern South America, CASA
front migration in the northern Andes: constraints from low-temperature GPS observations. J. South Am. Earth Sci. 15 (2), 157e171. http://dx.doi.org/
thermochronology in the Eastern Cordillera, Colombia. Tectonics 28, TC4004. 10.1016/S0895-9811(02)00018-4.
http://dx.doi.org/10.1029/2008TC002423. Wessel, P., Smith, W.H.F., 1998. New, improved version of the generic mapping tools
Parra, M., Mora, A., Lopez, C., Rojas, L.E., Horton, B.K., 2012. Detecting earliest released [abs.]. Eos (Trans., Am. Geophys. Union) 79, 579.
shortening and deformation advance in thrust belt hinterlands: example from Wittmann, H., von Blanckenburg, F., Guyot, J.L., Laraque, A., Bernal, C., Kubik, P.W.,
the Colombian Andes. Geology 40 (2), 175e178. http://dx.doi.org/10.1130/ 2011. Sediment production and transport from in situ-produced cosmogenic
10
G32519.1. Be and river loads in the Napo River basin, an upper Amazon tributary of
Pedoja, K., 2003. Les terrasses marines de la marge Nord Andine (Equateur et Nord Ecuador and Peru. J. South Am. Earth Sci. 31 (1), 45e53. http://dx.doi.org/
Pe0 rou): Relations avec le contexte ge0 odynamique (Ph.D. thesis). Univ. Paris VI, 10.1016/j.jsames.2010.09.004.
Paris, p. 413. Wijninga, V.M., 1996. Palynology and Paleobotany of Neogene Sediments from the
Platt, J.P., 1993. Mechanics of oblique convergence. J. Geophys. Res. 98, 16,239e High Plain of Bogotá (Colombia): Evolution of the Andean Flora from an
16,256. Ecological Perspective (PhD thesis). University of Amsterdam, Amsterdam,
Plumb, R.A., 1982. Breakouts in the geothermal well, Auburn, N.Y [abs.]. Eos (Trans., p. 370.
Am. Geophys. Union) 63, 1118. Younes, A., Engelder, T., 1999. Fringe cracks: key structures for the interpretation of
Plumb, R.A., Hickman, S.H., 1985. Stress-induced borehole enlargement: a com- progressive Alleghanian deformation of the Appalachian Plateau. Geol. Soc. Am.
parison between the four-arm dipmeter and the borehole televiewer in the Bull. 111, 219e239.
Auburn geothermal well. J. Geophys. Res. 90 (B7), 5513e5521. http://dx.doi.org/ Zoback, M.D., Moos, D., Mastin, L.G., Anderson, R.N., 1985. Well bore breakouts and
10.1029/JB090iB07p05513. in situ stress. J. Geophys. Res. 90, 5523e5530.
Pulido, N., 2003. Seismotectonics of the northern Andes (Colombia) and the Zoback, M.L., Zoback, M.D., 1989. Tectonic stress field of the continental U.S. In:
development of seismic networks. Bull. Int. Inst. Seismol. Earthq. Eng., 69e76. Pakiser, L., Mooney, W. (Eds.), Geophysical Framework of the Continental United
Special Edition. States. Geol. Soc. Am. Mem. 172, 523e539.

You might also like