6

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

TECTONICS, VOL. 31, TC3008, doi:10.

1029/2011TC003089, 2012

Kinematic evolution of Andean fold-thrust structures along


the boundary between the Eastern Cordillera and
Middle Magdalena Valley basin, Colombia
Javier Sánchez,1 Brian K. Horton,1,2 Eliseo Tesón,3 Andrés Mora,3 Richard A. Ketcham,1
and Daniel F. Stockli1
Received 21 December 2011; revised 1 March 2012; accepted 20 March 2012; published 17 May 2012.
[1] Surface and subsurface data support a kinematic reconstruction of Cenozoic fold-thrust
deformation along the Eastern Cordillera-Magdalena Valley transition in Colombia. The
La Salina fault (LSF) marks the boundary between west-vergent Eastern Cordillera
structures and hinterland deposits of the Middle Magdalena Valley basin. Apatite fission
track and (U-Th)/He thermochronological results for the west-directed LSF reveal initial
hanging wall exhumation during middle Eocene–early Oligocene (45–30 Ma)
shortening, renewed exhumation in the early middle Miocene (18–12 Ma), and
accelerated late Miocene-Pliocene (12–3 Ma) exhumation. Vitrinite reflectance data
suggest maximum burial of 4–6 km, helping constrain Cenozoic basin architecture.
Mapping of the LSF reveals hanging wall Cretaceous–Eocene rocks in a broad
anticline-syncline pair with limited faulting and footwall Eocene–Quaternary basin fill in a
complex series of tight thrust-related folds. Limited displacement along the westernmost
(frontal) thrust suggests that shortening is largely accommodated by east-directed thrusting
within a broader triangle zone of a passive-roof duplex (and probable minor strike-slip
deformation). In the preferred kinematic restoration, the most recent phase of shortening to
transpressional deformation represents out-of-sequence reactivation of the LSF
consistent with irregular crosscutting relationships among footwall structures. Earliest
exhumation by 45–30 Ma in the Eastern Cordillera fold-thrust belt is correlated with
increased sedimentary lithic fragments and high compositional maturity in sandstones of
the adjacent Magdalena Valley basin. Exhumation since 15 Ma coincided with
decreased compositional maturity and elevated accumulation rates for the Real Group.
The compositional provenance shifts are attributed to westward advance of fold-thrust
deformation into the proximal (eastern) segments of the Magdalena Valley basin.
Citation: Sánchez, J., B. K. Horton, E. Tesón, A. Mora, R. A. Ketcham, and D. F. Stockli (2012), Kinematic evolution of
Andean fold-thrust structures along the boundary between the Eastern Cordillera and Middle Magdalena Valley basin, Colombia,
Tectonics, 31, TC3008, doi:10.1029/2011TC003089.

1. Introduction timing and magnitude of exhumation [Shagam et al., 1984;


Toro, 1990; Mora et al., 2006; Parra et al., 2009a; Horton
[2] Although previous efforts have addressed deformation
et al., 2010a]. The adjacent hinterland basin to the west, the
of the Eastern Cordillera (EC) fold-thrust belt in Colombia, Middle Magdalena Valley basin (MMVB), contains an
disagreement persists over potential relationships between
important record of EC exhumation. Many studies have pro-
the generation and growth of Andean structures and the
posed exhumation scenarios based on sedimentologic, petro-
1
graphic, and isotopic analyses of basin fill [Gómez et al.,
Department of Geological Sciences, Jackson School of Geosciences, 2005a; Caballero, 2010; Caballero et al., 2010; Nie et al.,
University of Texas at Austin, Austin, Texas, USA.
2
Institute for Geophysics, Jackson School of Geosciences, University of 2010, 2012; Moreno et al., 2011; Saylor et al., 2011]. Other
Texas at Austin, Austin, Texas, USA. investigations have used field mapping and structural analyses
3
Instituto Colombiano del Petróleo, Ecopetrol, Bucaramanga, along the EC-MMVB boundary to establish relationships
Colombia. between structures and sedimentary units [Acosta et al., 2004;
Corresponding author: B. K. Horton, Department of Geological Restrepo-Pace et al., 2004; Gómez et al., 2005a, 2005b; Sassi
Sciences, Jackson School of Geosciences, University of Texas at Austin, et al., 2007]. More recent approaches have demonstrated the
1 University Station C9000, Austin, TX 78712, USA. utility of low-temperature thermochronology of both bedrock
(horton@jsg.utexas.edu) and detrital materials in assessing exhumation of the axial to
Copyright 2012 by the American Geophysical Union. easternmost EC [Parra et al., 2009b; Horton et al., 2010b;
0278-7407/12/2011TC003089

TC3008 1 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 1. (a) Regional tectonic map of northwestern South America (BSF: Bucaramanga-Santa Marta
fault; PCB: Panama-Choco block). (b) Map of the Colombian Andes showing topography (color shading),
tectonic provinces (WC: Western Cordillera; CC: Central Cordillera; MMVB: Middle Magdalena Valley
basin; EC: Eastern Cordillera) and major structures (AA: Arcabuco anticline; BF: Boyaca fault; BP:
Bogota plateau; GF: Guaycaramo fault; GS: Guaduas syncline; IF: Ibague fault; LB: Llanos basin; LSF:
La Salina fault; NS: Nuevo Mundo syncline; PF: Palestina fault; SeF: Servita fault; SM: Santander massif;
SoF: Soapaga fault; SuF: Suarez fault; U.S.: Usme syncline; VA: Villeta anticlinorium).

Mora et al., 2010a; Moretti et al., 2010; Bande et al., 2012; 2006, 2010b]. Further critical to understanding construc-
Saylor et al., 2012]. tion of the EC are thermochronological exhumation records
[3] Integration of these techniques proves beneficial for [Parra et al., 2009b; Restrepo-Moreno et al., 2009; Mora
assessing not only structural style, fault timing, and defor- et al., 2010a; Saylor et al., 2012] along with results on
mation rates, but also evolving basin configuration during the evolving architecture of Cenozoic sedimentary basins
shortening, uplift, and exhumation. For EC fold-thrust [Corredor, 2003; Gómez et al., 2005a, 2005b; Bayona
deformation, different workers have interpreted variable et al., 2008; Horton et al., 2010a; Saylor et al., 2011;
influence of preexisting Mesozoic normal faults [Casero Bande et al., 2012].
et al., 1997; Branquet et al., 2002; Kammer and Sánchez, [4] The purpose of this study is to evaluate the structural
2006; Mora et al., 2009, 2010b], strike-slip structures transition in the northern Andes (Figure 1) from the W-
[Acosta et al., 2007; Sassi et al., 2007], and generating directed thrust system defining the W flank of the EC to the
contrasting estimates of 50 km to >200 km of total E-W hinterland fill of the MMVB in order to constrain the kine-
shortening [e.g., Dengo and Covey, 1993; Cooper et al., matic evolution of this segment of the fold-thrust belt and its
1995; Roeder and Chamberlain, 1995; Mora et al., 2006; connection to the exhumation history of the EC. Construc-
Sarmiento-Rojas et al., 2006]. The EC is also characterized tion of sequentially restored structural cross sections, based
by thin- and thick-skinned geometries, basement-involved on field mapping and interpretation of 2-D seismic data of
and basement-detached deformation, and contrasting fold- the eastern MMVB, help document the evolving geometry
thrust vergence with common backthrusting, tectonic wedg- of the fold-thrust belt. The 4300 km2 study area contains
ing, passive-roof duplexing, and triangle-zone deformation the western foothills of the EC and part of the adjacent
[Restrepo-Pace et al., 2004; Toro et al., 2004; Mora et al., MMVB (Figure 2). A substantial difference in elevation and

2 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 2. (a) Geological map and (b) generalized Jurassic–Pliocene stratigraphic column for the bound-
ary between the Middle Magdalena Valley basin and Eastern Cordillera at 6–6.5 N (modified from Rolon
[2004]). Map shows location of 2D seismic lines and wells (white dots) used in the study relative to the
main structures (OA: Opon anticline; OS: Opon syncline; CF: Cimitarra fault).

relief between these two provinces defines the trace of the shortening during accretion of the Western Cordillera.
principal thrust structure, the E-dipping La Salina fault Studies of provenance, sedimentary facies, and limited
(LSF) (Figures 1 and 2). thermochronological data indicate exhumation and positive
topography in the Central Cordillera since latest Cretaceous
2. Regional Geologic Framework time [Gómez et al., 2003; Parra et al., 2009a].
[8] 3. The Eastern Cordillera (EC) is a bivergent fold-
[5] Tectonic interactions among the South American, thrust belt about 200 km wide [Dengo and Covey, 1993;
Caribbean, and Nazca plates, including subduction, obduc- Cooper et al., 1995] that overprints a broad rift system
tion, terrane accretion, and modern oblique transpression, developed during Jurassic-Cretaceous extension [Sarmiento,
controlled the Cenozoic evolution of the northern Andes (see 2001; Sarmiento-Rojas et al., 2006]. In the hinterland
Figure S1 in the auxiliary material) [Etayo-Serna et al., between the CC and WC, the Magdalena Valley of Colombia
1983; McCourt et al., 1984; Kellogg, 1984; Duque-Caro, is a NNE-trending intermontane basin (Figure 1b) subdivided
1990; Trenkamp et al., 2002].1 The northern Andean oro- into the Lower, Middle and Upper Magdalena Valley. The
gen is composed of three NNE-trending ranges separated by Middle Magdalena Valley basin (MMVB), the focus of this
intermontane hinterland basins (Figure 1a). study, is bounded to the west by poorly known thrust struc-
[6] 1. The Western Cordillera (WC) contains oceanic tures and the Palestina fault and to the south by the Ibague
fragments of Pacific affinity accreted to South America dur- right-lateral strike-slip fault [Cediel et al., 2003; Montes
ing the Paleogene [Cooper et al., 1995; Cediel et al., 2003]. et al., 2003, 2005; Acosta et al., 2007]. The eastern basin
These terranes may be derived in part from arc equivalents of margin is defined by the EC-bounding LSF and, farther
the Antilles Caribbean plateau [Pindell and Tabbutt, 1995; north, the Bucaramanga-Santa Marta left-lateral fault system
Spikings et al., 2001, 2005; Mann et al., 2006]. bounding the Santander massif (Figure 1).
[7] 2. The Central Cordillera (CC), a Mesozoic-Cenozoic [9] Basin fill of the MMVB (Figures 2b and 3) consists of
magmatic arc, was affected by latest Cretaceous-Paleogene Jurassic–Lower Cretaceous (Berriasian) fluvial clastic/vol-
canic rocks (Girón and Arcabuco-Los Santos Formations),
1
Auxiliary materials are available in the HTML. doi:10.1029/ Lower Cretaceous calcareous and siliciclastic shallow-
2011TC003089. marine strata (Basal Lime Group: Cumbre, Rosablanca, Paja,

3 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 3. Jurassic–Pliocene time-stratigraphic chart showing different lithostratigraphic units of the


study region, including characteristic lithologies and depositional environments. Samples of this study
are displayed relative to the main structures (AA: Arcabuco anticline; BF: Boyaca fault; CA: Cobardes
anticline; LSF: La Salina fault; PF: Palestina fault; SuF: Suarez fault). Stratigraphic data compiled from
Cooper et al. [1995], Sarmiento [2001], Cediel et al. [2003], Bayona et al. [2003, 2008], Rolon [2004],
Gómez et al. [2005a, 2005b], Caballero et al. [2010], and Mora et al. [2010a].

and Tablazo Formations), Albian–Maastrichtian shallow- of surface uplift associated with basin evolution remains
to-deep marine strata (Simití and La Luna Formations), and elusive [e.g., Saylor et al., 2009].
Upper Cretaceous–Paleocene shallow-marine to nonmarine [10] The Magdalena Valley basin has evolved in connec-
siliciclastic rocks (Umir and Lisama Formations) [Villamil, tion with the eastward advance of Andean orogenesis that
1999; Rolon, 2004]. In the southern MMVB, coarser facies began in the Late Cretaceous. The MMVB succession con-
in the uppermost Cretaceous-Paleocene succession are asso- sists of an eastward-thickening wedge with 2–10 km of
ciated with fluvial to alluvial fan deposition (Cimarrona, Seca, Mesozoic-Cenozoic strata deposited on Proterozoic to lower
and Hoyon Formations) [Gómez et al., 2003]. Cenozoic flu- Paleozoic basement [Cediel et al., 2003]. Along the W
vial/lacustrine rocks (La Paz, Esmeraldas, Mugrosa, Col- margin, a regional angular unconformity recorded a major
orado, and Real Formations) deposited during construction period of late Paleocene-early Eocene erosion [Schamel,
of the CC and EC [Dengo and Covey, 1993; Cooper et al., 1991; Rolon, 2004; Gómez et al., 2003, 2005a; Horton,
1995; Rolon, 2004; Nie et al., 2010; Moreno et al., 2011; 2012]. This depositional gap merges with conformable
Horton, 2012] consist of interbedded sandstone, conglom- strata in the eastern part of the basin [Pardo-Trujillo et al.,
erate, and mudstone [Gómez et al., 2003, 2005a]. Without 2003], recording exhumation due to eastward advance of
constraints on paleoelevation through time, the magnitude deformation from the Central Cordillera to the eastern

4 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

MMVB (Figure 3). Additional minor unconformities record by Ecopetrol. Key interpreted horizons include Upper
the local growth of several structures in the eastern MMVB Cretaceous to upper Miocene units, including the Umir,
[Gómez et al., 2003, 2005a; Restrepo-Pace et al., 2004; Lisama, La Paz, Esmeraldas, Mugrosa, Colorado, and Real
Parra et al., 2012]. Formations.
[14] Four cross sections constructed on seismic profiles
3. Structural Geology were subjected to time-depth conversions using velocities
provided by check shot logs from wells and root mean
3.1. Structural Background squared (RMS) migration velocities from seismic profiles
[11] Diverse interpretations of EC structural styles have (see Table S1 in the auxiliary material). These results were
resulted in discrepancies on the geometry, kinematic history, then integrated with surface data to generate structural sec-
and total shortening of this bivergent orogen. Basement- tions. Geometric restoration using line-length balancing
involved faulting of several structural highs has been asso- techniques [e.g., Dahlstrom, 1969; Hossack, 1979] and a
ciated with inversion of the principal normal faults of a fault-parallel flow algorithm [Suppe, 1983] enabled evalua-
Mesozoic rift basin, including the main reverse fault systems tion of several admissible sections in order to obtain the
along both the W and E margins of the EC [Colletta et al., most appropriate structural model.
1990; Cooper et al., 1995; Mora et al., 2006, 2008]. Alter-
natively, several detachment levels within the Cretaceous 3.3. Surface and Subsurface Results
section support a thin-skinned style of shortening [Dengo [15] The W-directed LSF separates the MMVB footwall
and Covey, 1993; Corredor, 2003; Restrepo-Pace et al., from the EC hanging wall (Figure 2). Whereas the LSF
2004; Cortés et al., 2006]. Erosion may further influence footwall is characterized by a complex series of tight fault-
the morphology and kinematics of thrust belts [Horton, related folds, the hanging wall displays gentler folding and
1999; Hilley and Strecker, 2004; Mora et al., 2008]. The little surface expression of significant faulting (Figure S3).
competing models differ principally in the geometry and The footwall exposes steep (50–80 ) to overturned Cenozoic
magnitude of shortening attributed to the various structures strata in the faulted Opon anticline and tightly overturned
(Figure S2). Dengo and Covey [1993] proposed thin-skinned Opon syncline, both of which plunge 10 NE. Farther
Miocene thrusting overprinted by Pliocene basement- south, a footwall syncline with a steep E flank (50–70 )
involved faulting, with 150 km of total shortening. In exposes the Paleocene Lisama Formation to upper Miocene
contrast, others suggest that thin-skinned deformation was a Real Group. The immediate hanging wall consists of the
late Miocene and younger process along the outer flanks of symmetric De Armas syncline with 30 limbs of Paleocene-
the EC adjacent to the major reverse-fault inversion struc- Eocene strata. However, the broader EC is dominated by
tures, resulting in total shortening of 58 to 105 km a thick Jurassic–Cretaceous section exposed in two large-
[Colletta et al., 1990; Cooper et al., 1995; Cortés et al., wavelength structures. The Los Cobardes and Portones
2006; Mora et al., 2006, 2010b; E. Tesón et al., Relation- anticlines, plunging to the south and north, respectively,
ship of Mesozoic graben development, stress, shortening show an en echelon relay map pattern (Figure 2) potentially
magnitude, and structural style in the Eastern Cordillera of associated with an inherited rift configuration [Schamel, 1991;
the Colombian Andes, submitted to Thick-Skin-Dominated Bayona et al., 2003; Kammer and Sánchez, 2006; Mora
Orogens: From Initial Inversion to Full Accretion, edited by et al., 2006].
M. Nemcok, A. Mora, and J. Cosgrove, Geological Society [16] Subsurface mapping of the LSF suggests an E-dipping
Special Publication, 2011]. Roeder and Chamberlain [1995] hanging wall ramp linked to a lower flat beneath the De
formulated an extreme end-member model based on litho- Armas syncline (Figure 4). Although seismic resolution is
spheric-scale ramp-flat thrusting to obtain 230 km of limited, a deeper ramp is interpreted to cut basement. The
shortening. symmetric De Armas syncline is considered the product of
[12] Along the EC-MMVB boundary, a broad negative W-flank uplift along the W-directed LSF and E-flank uplift
gravity anomaly is associated with crustal thickening in the as part of a fault-bend fold above a deeper flat-ramp geom-
EC adjacent to 10 km-thick fill of the MMVB [Toro, etry, in addition to E-directed backthrusts detached in shal-
1990]. Additional gravity models for the westernmost EC low Cretaceous levels (Figure 4).
show asymmetric anomalies potentially representing a relict [17] In the LSF footwall, subsurface data show the tight
Jurassic basin and basement high along with a more- Opon syncline flanked by the main strand of the LSF and a
complete Cretaceous section linked to possible thrust repe- footwall triangle zone (Figure 4a). The triangle zone is
tition by the LSF [Dengo and Covey, 1993]. These models constituted by a W-directed splay of the LSF cut by an E-
indicate a basement-involved structural style for the W flank directed back thrust that may form the carapace of a broader
of the EC, which could coexist with a thin-skinned style passive-roof duplex involving deeper Cretaceous and base-
where the Cretaceous cover is thicker. ment rocks. Additional footwall structures are identified to
the south and west (Figures 2 and 4). To the south, the
3.2. Methods oblique, NW-striking, SW-directed Cimitarra fault repre-
[13] Targeted field mapping and documentation of sents the southern termination of the tight Opon anticline
mesoscale structures (Figures 1 and 2) was performed using and syncline structures. To the west, the buried “frontal”
existing geologic maps and radar images with 30 m spatial thrust defines the westernmost W-directed structure in the
resolution. Subsurface structure was assessed through MMVB. Both footwall structures appear to exhibit Creta-
mapping and construction of cross sections using 49 2-D ceous detachment levels to the north and ramp geometries to
seismic reflection profiles and data from 8 wells provided the south.

5 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 4

6 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

[18] Four seismic sections were converted to depth to define lacks a flat in the Cretaceous section and consists of a ramp
spatial variations in subsurface geometries (Figures 4a, 4b, cutting basement. In the footwall, the steep E flank of a
S4a, and S4b). The characteristic structures in the area are best syncline is controlled by a passive-roof duplex; to the west,
expressed along the two central sections (Figures 4a and 4b). the frontal thrust has a limited displacement and generates a
The main strand of the LSF generally shows 4 dip segments subtle anticline.
for the 3 northern sections: an upper segment dipping 35 , a [23] The Los Cobardes and Portones anticlines in the LSF
<10 detachment, a 30 dip through the Cretaceous section, hanging wall are cored by Jurassic rocks, suggesting either a
and a lower 40 dip segment through the pre-Cretaceous Jurassic detachment level, which would require substantial
section. In contrast, the southern section shows a simpler listric shortening, or basement-involved faulting suggestive of
geometry revealed by decreasing dip (55 to 30 ) with depth. limited shortening and possible inversion of Mesozoic rift
structures [Cooper et al., 1995; Sarmiento, 2001; Mora et al.,
3.4. Cross Sections and Restorations 2006, 2008]. Evidence for inversion includes: (1) coarse-
[19] Cross sections constructed from field and subsurface grained, fault-proximal alluvial-fan and fluvial/lacustrine
data (Figures 5a, 5b, S5a, and S5b) show the W-directed facies of the Jurassic Girón Formation [Sarmiento-Rojas
LSF accommodating significant displacement (16–17 km). et al., 2006; Kammer and Sánchez, 2006]; (2) a relay
From west to east, the principal structures include the frontal zone between the Los Cobardes and Portones anticlines
thrust, the LSF system (including the footwall Opon anti- potentially due to an inherited en echelon basement rift
cline-syncline pair and hanging wall De Armas syncline), configuration; (3) correlation of high topography with
and the broad Los Cobardes and Portones anticlines. pre-Cretaceous exposures, indicating significant basement
[20] The frontal thrust cuts the Pliocene MMVB section uplift possibly facilitated by reactivation of pre-existing
but displays limited displacement (200–300 m), implying faults with low frictional resistance [Zhou et al., 2006].
that creation of significant structural relief and exhumation [24] Restoration of four line-length balanced cross sec-
of Cretaceous-Cenozoic fill along the MMVB-EC boundary tions (Figures 5 and S5) reveals similar shortening values of
was accommodated by some other mechanism. Possible 22–31 km (22–28%). Our cross sections suggest approxi-
solutions include (1) greater displacement along other buried mately E-W shortening accommodated by W-directed
W- or E-directed thrusts, (2) a large-scale fault-propagation translation on a forethrust (LSF) and kinematically linked E-
fold with shortening by thrust displacement at lower levels directed backthrusting (LB, UB). This tectonic wedging can
accommodated by folding in upper levels, or (3) a strike-slip be restored by considering an E-directed back thrust (or roof
to transpressional flower structure with diminished dis- thrust) as a passive fault that originates at the upper tip of an
placement on the most external splay faults. Our preferred active W-directed fault, creating an intact block bounded by
interpretation based on field and seismic observations is a two oppositely verging faults (Figure 6). This system
passive-roof duplex (or tectonic wedge) system beneath the evolves with the generation of a second back thrust
LSF, where shortening at depth along the main W-directed (Figure 6d) and finally an out-of-sequence forethrust
thrust system is taken up by E-directed backthrusting near (Figure 6f), resulting in a triangle zone geometry. Support
the surface. This interpretation consists of two backthrusts: for this interpretation comes from irregular crosscutting
an upper back thrust (UB) detached in the Oligocene section relationships among the LSF, some footwall structures, and
and a lower back thrust (LB) in the Upper Cretaceous sec- upper Miocene-Quaternary basin fill (Figure 2a), suggesting
tion, with the UB feeding slip into the Opon anticline-syn- out-of-sequence deformation and complex westward
cline pair (Figures 5a and S5a). Nevertheless, we emphasize advance of the deformation front.
that a potentially significant degree of oblique shortening is
possible for selected areas, including cases such as the
transverse NW-striking Cimitarra fault in the LSF footwall 4. Thermochronology
(Figure 2a), which may explain the modest inconsistency in [25] Recent thermochronological and geochronological
area conservation (Figure 5b). results better define the timing of EC deformation. New fission
[21] The LSF hanging wall consists of a relatively intact track data for apatite (AFT) and zircon (ZFT), and U-Th/He
EC block with structural relief substantially higher (by results for apatite (AHe) and zircon (ZHe) help link exhuma-
12 km) than the adjacent MMVB footwall. The De tion in the fold-thrust belt to thrusting in the western EC and
Armas syncline (AS) is situated between the LSF and E- proximal (eastern) MMVB. New and published vitrinite
directed backthrusts mainly detached in the Upper Creta- reflectance (Ro) data provide estimates of maximum paleo-
ceous Umir and La Luna Formations. To the east, the temperature and burial depth that further help constrain the
regionally continuous Los Cobardes anticline (CA) shows generation of shortening-related structures in the EC thrust-
modest asymmetry in which its E limb is steeper than the belt and MMVB hinterland.
W limb, possibly suggestive of east vergence.
[22] For the southern cross section (Figure S5b), south- 4.1. Methods
ward termination of the hanging wall De Armas syncline [26] New results consist of 20 AFT, 8 AHe, 5 zircon ZHe,
(Figure 2) suggests a geometric change in which the LSF 2 ZFT, and 7 Ro analyses across the LSF region. Several

Figure 4. Structural cross sections and corresponding seismic data for (a) N-Opon and (b) S-Opon sections (locations
shown in Figure 2). Sections integrate surface and subsurface information to illustrate the main features: principal fault dis-
placement (16–17 km) along the LSF; minor deformation along the frontal thrust (FT) accommodated by a passive roof
duplex in the LSF footwall; basement-involved deformation; and possible inverted structures (Los Cobardes (CA) and Por-
tones (PA) anticlines) to the east. Strata color scheme matches Figure 2b.

7 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 5. Restoration of the structural cross sections (Figure 4) for (a) N-Opon and (b) S-Opon sections,
including values of horizontal shortening. The S-Opon section (b) shows an inconsistency in area conser-
vation (pink area) possibly associated with oblique shortening along the transverse Cimitarra fault
(Figure 2a).

samples were analyzed using multiple techniques (Figure 7), suggests multiple populations. The age dispersion of a multi-
thus enhancing the exhumation record for particular locali- population sample indicates that the annealing properties of
ties and horizons. Fission track, (U-Th)/He, and Ro analyses the grain were not sufficiently homogeneous or there was
were performed at Apatite to Zircon Inc., the University of differing inheritance. Because pooled ages of multipopulation
Kansas, and the Ecopetrol, respectively. samples are not reliable, in such cases, sample grains were
[27] AFT and ZFT results were generated using the LA- divided into separate populations based on similarities in the
ICP-MS method [Donelick et al., 2005], then classified as etching parameter Dpar. For each sample, the age of the pop-
single- or multikinetic population based on their individual ulation passing the c2 test that encompasses the maximum
grain age distributions according to the chi-square (c2) test: number of grains is considered to represent the sample age.
P[c2] > 5% supports a single population, but a P[c2] < 5%

8 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

[28] AHe and ZHe analyses were conducted on 4 grains


per sample and are reported as sample mean ages with 1s
standard deviation. We discarded a small number of analyses
with a relatively low alpha ejection correction (Ft < 0.6)
[Farley et al., 1996] or an age inconsistent with the geo-
logical context or AFT cooling constraints.
[29] Ro analyses on randomly oriented vitrinite (Rorand)
were converted to maximum reflectance values (Romax)
[Zhang and Davis, 1993] then to paleotemperatures [Sweeney
and Burnham, 1990] assuming a normal heating rate of
1 C/Myr. Then, based on a geothermal gradient calculated
from paleotemperature trends, maximum burial depths were
obtained and integrated with thermal modeling and kine-
matic restorations.
[30] Past positions of the partial annealing zone (PAZ) for
AFT and partial retention zone (PRZ) for AHe can be
inferred by comparing cooling and depositional ages, eval-
uating the burial depth based on Ro data, reviewing AFT
track-length variations and age distributions relative to
stratigraphic position [Brandon et al., 1998; O’Sullivan and
Wallace, 2002], as well as thermal modeling [Ketcham et al.,
2007]. It is possible to identify: (1) thermally reset samples
(maximum temperature exceeding the total annealing or
diffusion temperature for AFT or AHe, respectively), which
record the last major exhumation event from beneath the
PAZ or PRZ; (2) partially reset samples (slightly cooler or
equal to the total annealing or diffusion temperature), which
are most informative for thermal history inversion modeling;
and (3) non-reset samples (cooler than the lower temperature
boundaries of the PAZ or PRZ), which reflect detrital
provenance and/or sediment source information.
[31] To establish exhumation patterns, 8 samples were
modeled with the AFT annealing model of Ketcham et al.
[2007], AHe diffusivity model of Flowers et al. [2009],
and ZHe diffusion parameters of Reiners et al. [2004] using
HeFTy software [Ketcham, 2005]. Model results enable us
to interpret the ages provided by the partially reset samples
and develop a thermal history that integrates different ther-
mochronometers and Ro data.
4.2. Results
4.2.1. Vitrinite Reflectance (Ro)
[32] LSF hanging wall samples from Lower Cretaceous–
Paleocene units show increased Ro index, maximum paleo-
temperature, and burial depth with lower stratigraphic positions.
The geothermal gradient calculated for hanging wall units
(Lisama, Umir, La Luna) is 30 C/km, with paleotemperatures
of 120–200 C and maximum burial depths of 4–6 km. A
Lower Cretaceous (Paja Formation) sample departs consid-
erably from Upper Cretaceous–Paleocene samples, suggest-
ing the sample may have been in a different block that
underwent less burial. Ro data from MMVB well samples of
Eocene La Paz to Miocene Colorado in the LSF footwall
enable calculation of a paleotemperature of 50–115 C, a
geothermal gradient of 25 C/km (using the linear paleo-
temperature:thickness relationship); and maximum burial
depth of 1.2–3.7 km (Table S2 and Figure S6).
Figure 6. (a–f) Simple forward model depicting the process 4.2.2. Low-Temperature Thermochronology
of tectonic wedging, where most shortening is accommo- [33] LSF hanging wall samples generally show cooling
dated by passive backthrusts (including possible generation ages significantly younger than depositional ages (Figure 8
of passive-roof duplex structures), ending with later develop- and Tables 1, 2, and 3), indicating fully reset samples situ-
ment of a triangle zone due to out-of-sequence thrusting. ated beneath the isotherms of total annealing and total

9 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 7. Geologic map showing location of thermochronological and vitrinite reflectance samples
along the boundary between the Middle Magdalena Valley basin and Eastern Cordillera. Enlarged map
area shows color- and symbol-coded samples for vitrinite, and AHe and ZHe thermochronology, and sand-
stone petrography.

diffusion (for AFT and AHe, respectively) prior to cooling. section (Colorado, Real, and Mesa) show ages older than
Six samples (15, 14, QA-05-19, 28, 29, 30) have AFT ages depositional ages, consistent with non-reset samples. Accor-
between 20 and 5 Ma, which can be associated with one or ding to the Ro paleotemperatures, the PAZ for AFT should
more Neogene cooling episodes. One sample (31) from the be in the Oligocene-Eocene section. AHe cooling ages of
base of the Eocene La Paz Formation shows a cooling age of footwall samples appear younger than depositional ages
35 Ma close to the depositional age. This indicates a par- below the middle Real Group, providing evidence of <6 Ma
tially reset AFT age and a PAZ in the hanging wall, con- cooling, consistent with the hanging wall cooling record.
sistent with Ro data for a sample (27) at this level indicating AHe cooling ages above the middle Real approach deposi-
a maximum paleotemperature of 110 C. This suggestion tional ages, which may indicate a PRZ in the Real Group
of an early cooling event may be explored with thermal situated above the footwall PAZ.
modeling. Overall, hanging wall cooling ages reported by [35] The high dispersion of the Miocene Real Dpar values
AFT and AHe are relatively close to each other (2–8 Myr (Figure 8) could indicate more than one source for apatite
age difference), implying relatively rapid cooling. grains. Beneath the Miocene section, Dpar values are gener-
[34] LSF footwall samples (17, 18) show AFT cooling ages ally smaller, indicating a possible difference in source rock.
for the Eocene Esmeraldas and Oligocene Mugrosa Forma- Fast and slow annealing apatites (≤1.75 mm and ≥1.75 mm,
tions close to depositional ages (Figure 8 and Table 1), sug- respectively) within the Upper Cretaceous–Oligocene section
gesting these samples may be partially reset and represent a may cause the dispersion in AFT ages; thus, thermal model-
PAZ. Six samples (16, 20, 22, 23, 24, 26) for the overlying ing is required to better constrain cooling. AFT mean track

10 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 8. Plot of AHe ages, AFT ages, and corresponding Dpar and track length data according to strati-
graphic level in the footwall and hanging wall of the La Salina fault (LSF). The relationships between
depositional ages (dotted line) and thermochronometer ages suggest a possible partial annealing zone
(PAZ) for AFT in both blocks and a partial retention zone (PRZ) for AHe in the footwall. Elevated Dpar
values in the Miocene section may indicate a different provenance (see text for details). Mean track length
data show an upsection increase in the footwall, consistent with interpretation of a basal PAZ.

lengths increase upsection, consistent with a transition from a depths below the AFT total annealing isotherm. This reset
PAZ in the Oligocene-Eocene section to an overlying non- sample shows initial cooling at 12–6 Ma and rapid cooling of
reset interval. 13–25 C/Myr. Umir sample 28 (Figure 10b) shows an onset of
[36] ZHe results for the LSF hanging wall reveal cooling exhumation at 15–7 Ma and a cooling rate of 8–17 C/Myr.
ages significantly older than depositional ages, indicating Paleocene Lisama sample 29 (Figure 10d) from the E flank of
these are non-reset samples bearing detrital ages (Figure 9 the De Armas syncline indicates a reset AFT sample with 23–
and Table 3). Similarly, Ro values show maximum paleo- 8 Ma cooling at 5–12 C/Myr. A possibly partially reset
temperatures less than ZHe closure temperatures (180 C) Paleocene Lisama sample 31 (Figure 10c) from the W flank of
and ZFT data (sample 112) show cooling ages older than the De Armas syncline shows initial AFT cooling at 43–35 Ma
depositional ages (Table S3). Results for a Jurassic sand- and slow cooling of 2–3 C/Myr.
stone (sample 114) with poor apatite recovery tentatively [39] Among footwall samples, Oligocene Mugrosa sam-
suggest an AFT cooling age of 6.7  1.6 Ma for the ple 17 (Figure 10e) from the E flank of the Opon syncline
Cobardes anticline (Table 1). was exhumed from depths above the AFT total annealing
4.2.3. Thermal Modeling isotherm; this partially reset sample records initial cooling
[37] Four samples from the LSF hanging wall were modeled at 18–10 Ma at a rate of 5–8 C/Myr. Eocene Esmeraldas
integrating AFT results (ages, track lengths, Dpar values), Formation sample 18 (Figure 10f) along the W flank of the
AHe, ZHe, and Ro information, according to the available data Opon syncline is partially reset and shows 11–3 Ma cooling
for each sample. Additional geological constraints include the at 6–20 C/Myr. Miocene Colorado sample 20 (Figure 10h)
sample’s depositional age, stratigraphic burial record (which from the W flank of the faulted Opon anticline indicates
delimits the heating rate, normally 1–5 C/Myr), a broad pre- slow cooling with possible earliest cooling constrained to
depositional history guided by the ZHe cooling age, and 16–3 Ma at a rate of 3–16 C/Myr. Results from Miocene
present surface temperature of 20 C (Figure 10 and Table S4). Colorado sample 26 (Figure 10g), from a different fault
[38] Among hanging wall samples, Upper Cretaceous Umir block (W flank of Opon syncline) than sample 20, indicate
Formation sample 14 (Figure 10a) indicates exhumation from

11 of 24
TC3008

Table 1. Results for Apatite Fission Track (AFT) Thermochronology Analyzed With Laser Ablation-Inductively Coupled Plasma-Mass Spectrometry (LA-ICP-MS)a
Pooled Age
Area LAICPMS Track After
Number Stratigraphic Analyzed Fission Mean Constraining
d
of Longitude Latitude Elevation Age Dparb Dperc Ns (W) S (PW)e 1s S(PW) Track Length Dpar
43 238
Sample Grains ( W) ( N) (m) Unit (Ma) (mm) (mm) (tracks) (cm2) (cm2) (cm2) x MSf 1sx MS Ca U P(g2)g Age(Ma) (mm) Populationsh

24 40 73.9371 6.3927 143 lower Mesa 4–5 2.31 0.64 202 1.09E-03 8.23E-01 2.64E-02 170,003 0.3161 3.34E+02 2.64E+02 0 20.8+/ 1.7 13.82 22.5+/ 6.1
16 40 73.8121 6.4605 314 middle Real 7–10 2.4 0.65 264 7.90E-04 4.74E-01 1.28E-02 159,333 0.3 2.60E+02 1.06E+02 0.0001 44.2+/ 3.1 12.04 48.9+/ 4.3
22 40 73.9218 6.3817 202 middle Real 7–10 2.25 0.62 463 1.34E-03 1.13E+00 3.24E-02 166,259 0.3104 3.06E+02 1.61E+03 0 34.1+/ 2.0 13.38 35.9+/ 2.2
23 40 73.9248 6.3813 195 middle Real 7–10 2.77 0.86 198 1.37E-03 8.47E-01 2.12E-02 168,131 0.3132 2.90E+02 1.66E+02 0 19.6+/ 1.5 14.23 13.8+/ 1.9
26 37 73.841818 6.451608 177 upper Colorado 15–17 1.74 0.41 110 4.70E-04 3.71E-01 1.03E-02 172,471 0.3431 8.52E+01 2.27E+01 0.0057 25.5+/ 2.6 13.53 25.4+/ 2.9
18 20 73.885163 6.384694 134 upper Esmeralda 30–33 1.76 0.39 62 1.95E-04 1.70E-01 4.71E-03 162,375 0.3046 2.60E+02 1.85E+02 0.7072 29.5+/ 3.9 11.28 29.5+/ 3.9
20 38 73.898721 6.383939 239 upper Colorado 15–17 1.91 0.42 381 4.24E-04 7.44E-01 1.45E-02 164,294 0.3075 2.79E+02 2.22E+02 0 41.9+/ 2.4 11.95 41.8+/ 7.9
17 39 73.864651 6.369856 393 lower Mugrosa 29–30 2.38 0.71 47 4.31E-04 1.36E-01 2.66E-03 161,018 0.3026 2.44E+02 3.50E+03 0.3129 27.8+/ 4.1 11.66 27.8+/ 4.1
29 39 73.784065 6.382611 1105 middle Lisama 57–60 1.57 0.51 71 5.72E-04 6.46E-01 1.99E-02 175,025 0.3486 1.16E+02 1.58E+03 0 9.61+/ 1.19 13.31 8.82+/ 1.23
30 40 73.789809 6.375733 1146 lower La Paz? 40–45 1.91 0.5 216 9.92E-04 9.09E-01 2.46E-02 176,227 0.3512 9.54E+01 2.57E+01 0.0007 20.9+/ 1.6 13.48 13.4+/ 1.7
31 40 73.859758 6.356064 650 upper Lisama? 40–45 2.16 0.6 154 1.21E-03 3.41E-01 5.99E-03 177,308 0.3536 1.19E+02 2.07E+01 0.1683 39.9+/ 3.4 11.8 39.9+/ 3.4
28 40 73.751994 6.400875 692 middle Umir 65–79 1.81 0.54 50 9.62E-04 3.24E-01 6.35E-03 173,823 0.346 9.45E+01 2.31E+01 0.5695 13.4+/ 1.9 13.96 13.4+/ 1.9
14 36 73.79371 6.250189 980 middle Umir 65–79 1.85 0.42 63 6.72E-04 9.03E-01 2.19E-02 156,245 0.2954 2.27E+02 7.72E+01 0.8163 5.45+/ 0.71 13.41 5.45+/ 0.71

12 of 24
QA-05-19 20 73 47′24.01 6 19′40.65 775 upper Umir 65–79 1.92 170 Published in Mora et al. [2010a] Published in Mora et al. 3.05E+01 14.1+/ 1.2 9.6 14.1+/ 1.2
[2010a]
15 25 73.761825 6.288113 631 middle Umir 69–73 1.73 0.49 8 9.22E-05 1.03E-01 4.38E-03 157,438 0.2972 2.13E+02 4.39E+01 1 6.12+/ 2.18 11.44 6.12+/ 2.18
27 3 73.826943 6.459811 189 lower Real 12–15 1.84 0.3 4 2.67E-05 2.39E-02 1.85E-03 173,177 0.3446 7.96E+01 6.64E+00 0.1707 14.4+/ 7.3 12.65 14.4+/ 7.3
13 1 73.862898 6.410917 161 middle Colorado 17–20 0 0 0 9.60E-06 2.15E-03 2.15E-04 155,332 0.294 1.65E+02 5.19E+00 n.a. 0.000+/ 53.699
21 2 73.906866 6.381971 1105 lower Real 15-Dec 1.92 0.43 2 1.47E-05 8.05E-03 6.78E-04 165,277 0.309 2.33E+02 1.77E+01 0.7856 20.5+/ 14.6 20.5+/ 14.6
19 5 73.901145 6.379346 980 upper Colorado 17–20 1.81 0.35 4 3.01E-05 3.48E-02 2.38E-03 163,241 0.3059 1.91E+02 1.61E+02 0.2312 9.37+/ 4.73 11.96 9.37+/ 4.73
25 1 73.873419 6.413896 153 middle Mugrosa 26–29 1.71 0.71 6 9.71E-06 3.84E-02 3.84E-03 171,885 0.3418 7.48E+01 5.02E-01 n.a. 13.4+/ 5.6 13.4
114 2 73.487076 6.272908 2043 Arcabuco/Giron 140–150 1.74 0.47 19 1.94E-05 2.01E-02 1.99E-04 0.029 6.7+/ 1.6 11.65 6.7+/ 1.6
a
All errors are 1s.
b
Dpar = mean maximum diameter of fission track etch figures parallel to the crystallographic c axis.
c
Dper = mean maximum diameter of fission track etch figures perpendicular to the crystallographic c axis.
d
Ns = number of spontaneous fission tracks counted over area W (area analyzed).
e
S(PW) = Sum of PW for all grains evaluated; P = (238U/43Ca) for an apatite grain; W = area over which NS and P are evaluated.
f
x MS = calibration factor based on LA-ICP-MS of fission track age standards.
SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA

g
Chi-square probability P(c2) > 5% are considered to represent a single population of ages.
h
Ages after dividing age population, in the cases of samples with P(c2) < 5%, using Dpar parameter.
TC3008
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Table 2. Results for (U-Th)/He in Apatite (AHe) Thermochronology


Longitude Latitude Elevation Stratigraphic Age 8% U Th Sm He Mass
Sample ( W) ( N) (m) Unit Age (Ma) (Ma) (Ma) (ppm) (ppm) (ppm) Th/U (ncc/mg) (mg) Fta
24 73.937143 6.39265 143 lower Mesa 4–5 5.2 0.3 9.6 24.9 61.7 2.6 0.26 1 0.57
5.3 0.3 8.2 15.4 57 1.9 0.24 2.5 0.68
6.8 0.9 17.4 81.1 57.7 4.7 0.8 2.1 0.64
8 0.5 6.6 15.5 57.5 2.4 0.32 2.8 0.69
16 73.812135 6.46049 314 middle Real 7–10 4.5 0.3 15.7 24.3 36.9 1.6 0.36 2.5 0.68
4.6 0.3 20.8 28 49 1.3 0.46 2.1 0.67
4.3 0.3 22.9 58.8 79.4 2.6 0.62 5.2 0.71
2 0.1 2.5 11.5 62 4.6 0.05 6.2 0.75
22 73.92178 6.381677 202 middle Real 7–10 10.4 1.9 124.9 1.9 108.4 0 6.15 8.8 0.8
9.3 0.6 3.8 2.6 25.4 0.7 0.16 2.6 0.7
9.9 0.6 6.6 2.7 18.5 0.4 0.27 2.3 0.69
6.3 0.4 3.5 4.5 60.2 1.3 0.12 2.6 0.67
23 73.924817 6.381271 195 middle Real 7–10 9.5 0.6 6.5 11.9 35.2 1.8 0.3 1.3 0.61
9.3 0.6 4.5 7 57.1 1.6 0.21 1.5 0.62
9.3 0.6 9.7 14.7 31.6 1.5 0.45 2.6 0.67
5.9 0.4 6.6 9.9 26.3 1.5 0.19 1.9 0.66
18 73.885163 6.384694 134 upper Esmeralda 30–33 n/a n/a 0.5 2 0.7 4.4 150.4 2 0.63
3.1 0.2 23.5 119 202.5 5.1 0.59 2.5 0.67
3.4 13.3 13.1 50.3 162.5 3.9 0.74 5.4 0.75
3.3 0.2 14.5 76.3 148 5.3 0.39 1.9 0.64
17 73.864651 6.369856 393 lower Mugrosa 29–30 4.9 0.3 3.3 15.2 37.7 4.6 0.12 1.6 0.63
2.1 0.1 12.8 51.2 37.9 4 0.17 1.2 0.59
9.9 0.6 4 17.8 82.9 4.4 0.32 2.8 0.68
5.4 0.3 3.8 20.1 64.5 5.2 0.18 2.7 0.68
29 73.784065 6.382611 1105 middle Lisama 57–60 8 0.5 12.7 20 137.3 1.6 0.48 1.2 0.59
6.6 0.4 14.9 35.6 64.3 2.4 0.56 2.5 0.67
7.1 0.4 5 7.6 27.5 1.5 0.18 2.1 0.66
6 0.4 6.3 22 83.9 3.5 0.26 2 0.65
30 73.789809 6.375733 1146 lower La Paz? 40–45 6.2 0.4 5.1 14.2 37.2 2.8 0.19 2.2 0.66
5.2 0.3 2.1 3.3 26.2 1.5 0.07 6.9 0.77
5.3 0.3 59.6 90 43 1.5 1.6 2.8 0.68
5.1 0.3 16 34.4 47.5 2.2 0.47 3.6 0.7
a
Alpha ejection correction [Farley, 2000].

insufficient burial to reach the lower isotherm of the AFT Suczek, 1979; Ingersoll et al., 1984; Garzanti et al., 2007].
PAZ and show poorly constrained cooling below 60 C. Secondary factors include climate, relief, transport mecha-
nism, depositional environment, diagenesis, and drainage
size [Folk, 1980; Ingersoll et al., 1993]. Both the W and E
5. Sandstone Petrography flanks of the MMVB were established as major sources of
sediment [Gómez et al., 2005a; Caballero, 2010; Caballero
[40] Detrital modal signatures are largely a function of the
et al., 2010; Moreno et al., 2011]. The western source is the
tectonic setting of sediment source regions [Dickinson and

Table 3. Results for (U-Th)/He in Zircon (ZHe) Thermochronology


Longitude Latitude Elevation Stratigraphic Age 8% U Th Sm He Mass
Sample ( W) ( N) (m) Unit Age (Ma) (Ma) (Ma) (ppm) (ppm) (ppm) Th/U (ncc/mg) (mg) Ft
29 73.784065 6.382611 1105 middle Lisama 57–60 2.2 0.2 5737.6 4027 1675.7 0.7 56.79 3.3 0.73
94.9 7.6 139.6 53.9 0.8 0.4 53 1.7 0.68
64.3 5.1 90.8 30.5 1.6 0.3 23.58 2.1 0.69
62.2 5 75.1 7.9 0.1 0.1 18.63 2.7 0.72
30 73.789809 6.375733 1146 lower La Paz? 40–45 38.3 3.1 547.5 149.9 0.5 0.3 94.99 6.5 0.79
24.2 1.9 424.4 113.5 0.7 0.3 47.05 8.3 0.8
84 6.7 221.9 75.2 0.6 0.3 82.91 4.8 0.76
51.9 4.2 80.2 50.4 1.4 0.6 19.81 6 0.77
31 73.859758 6.356064 650 upper Lisama? 40–45 472.1 37.8 16.3 10 0.4 0.6 42.37 30 0.86
78.4 6.3 108.9 36.3 0.6 0.3 43.11 29.3 0.86
91.8 7.3 81.6 28.5 0.5 0.4 38.74 42.3 0.88
46 3.7 137.7 47.9 0.8 0.3 31.73 25.1 0.86
28 73.751994 6.400875 692 middle Umir 65–79 192.8 15.4 359.9 160.8 8 0.4 297.69 3.1 0.71
61.5 4.9 179.5 80.9 5.5 0.5 50.64 6.1 0.77
198.1 15.9 205 92.1 5.1 0.4 188.23 5.4 0.77
18.4 1.5 215.3 107 5.2 0.5 18 4.9 0.76
14 73.79371 6.250189 980 middle Umir 65–79 195.5 15.6 54.5 25.3 2.4 0.5 52.73 12.8 0.82
488 39 103.5 50 0.8 0.5 251.62 8 0.8
352.1 28.2 85.7 42.6 3.1 0.5 151.33 13 0.81
175.5 14 157.2 28 0.7 0.2 126.5 9.5 0.81

13 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 9. (a) Topographic profile, (b) structural cross section, and (c) thermochronological transect
showing AHe, ZHe, AFT, and ZFT results and vitrinite reflectance (Ro) data (profile and sample locations
shown in Figures 2, 4a, and 7). Clustering of thermochronometer ages and their relationship to deposi-
tional ages (black line) show that AFT samples are reset in the LSF hanging wall and possibly partially
reset in most of the footwall, AHe ages are generally reset in the footwall but possibly partially reset in
the west (upper Real Group, Mesa Formation), and ZHe and ZFT ages are not reset and therefore represent
detrital ages.

CC, a range constructed since the Late Cretaceous [Villamil, Our provenance study aims to evaluate detrital signals that
1999; Gómez et al., 2005b] with magmatic arc character- indicate the initiation of EC-derived sedimentation and rel-
istics and minor metamorphic assemblages [McCourt et al., ative contribution from EC and CC sources.
1984]. Significant proportions of volcanic and possibly
metamorphic (mainly schist and metapelite) fragments are 5.1. Methods
expected in sediment from the CC. Any sedimentary cover in
[41] Standard petrographic thin sections were prepared
the CC (up to 7–13 km) was likely eroded during Paleocene-
from 34 medium-grained Cenozoic sandstone samples col-
early Eocene time [Gómez et al., 2005a]. In contrast, the
lected mainly from the LSF footwall (Figure 7). Each thin
eastern source area is the EC fold-thrust belt initially con-
section was injected with blue dye and stained for plagio-
structed during middle-late Eocene to early Oligocene
clase and potassium feldspar. Modal compositional data
shortening [Nie et al., 2010], potentially supplying recycled
were collected following the Gazzi-Dickinson method
sedimentary material from an extensive Mesozoic basin
[Ingersoll et al., 1984] in which at least 450 points per thin
system [Caballero, 2010; Caballero et al., 2010; Moreno et al., section were counted. Matrix was not counted and an
2011]. Basement highs in the MMVB (La Cira-Infantas
attempt was made to distinguish the original mineral from
high) and EC (Santander massif) could also be important
authigenic minerals. Possible diagenetic loss of feldspar is
sources of metamorphic fragments (schist, quarzite, gneiss). difficult to assess and cannot be reliably evaluated with this

14 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 10. HeFTy thermal modeling results incorporating AFT, Ro, and AHe data for samples from the
La Salina fault system. Hanging wall samples: (a, b, c) E flank of the De Armas syncline; (d) W flank of the
De Armas syncline. Footwall samples: (e) E flank of the Opon syncline; (f, h) W flank of the Opon syncline;
(g) W flank of the Opon anticline. Modeled time-temperature paths must pass through and connect succes-
sive time-temperatures constraints (boxes) with alphanumeric labels depicting the number of times each
path segment was halved to allow variations in heating or cooling rate (1–4) and the degree of change in
rates (G: gradual; I: intermediate; E: episodic). The inversion model for each sample shows the best fit solu-
tion (black line) and portions of time-temperature paths yielding “good” (purple) and “acceptable” (green)
fits. In addition, other parameters and outputs of the model are shown: measured and modeled AFT age,
track-length, AHe age and Ro% values, number of AFT ages and lengths, goodness of fit (GOF) between
measured and modeled data [Ketcham, 2005], and the age of the oldest fission track not fully annealed.

15 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Table 4. Sandstone Point Count Parametersa


Ternary Plot Variable 1 Variable 2 Variable 3
Q-F-Lf Q = Qm + Qp + Qpt F = P + K + Lvf Lf = Lvl + Lm + Lsh + C [Folk, 1980]
Qt-F-L Qt = Qm + Qp + Qpt + C F=P+K L = Lvf + Lvl + Lm + Lsh
Qm-F-Lt Qm F=P+K Lt = Lvf + Lvl + Lm + Lsh + C + Qp + Qpt
Lm-Lv-Ls Lm Lv = Lvf + Lvl Ls = Lsh + C
a
Q (quartz), Qm (monocrystalline quartz), Qp (polycrystalline quartz), Qpt (polycrystalline quartz with tectonic fabric), Qt (total quartz), F (feldspar),
P (plagioclase felspar), K (potassium feldpar), L (lithic fragments), Lv (volcanic lithic fragments), Lvf (felsic volcanic grains), Lvl (lathwork and vitric
volcanic grains), Lm (metamorphic lithic fragments), Ls (sedimentary lithic fragments), Lsh (shale/mudrock grains), C (chert grains), Lt (total lithic
fragments).

method [Milliken, 1988]. Petrographic counting parameters Relative to underlying units, the Real Group shows a sharp
defined for this study (Table 4) were used to generate the decrease in quartz accompanied by increased Lt, with similar
modal point-count data (Table S5). Using compositional proportions of metamorphic, volcanic, and sedimentary
data, the sandstone classifications and provenance interpreta- fragments (Lm33Lv34Ls33). Most Real samples, especially
tions for each unit were based on various ternary diagrams, in upper levels, are rich in amphibole. All samples show
including Q-F-Lf [Folk, 1980] along with Qt-F-L and Qm-F- similar amounts of Qp, which is slightly lower than under-
Lt plots [Dickinson and Suczek, 1979; Dickinson et al., 1983; lying units.
Garzanti et al., 2007]. [47] Samples of the Pliocene Mesa Formation (2 samples)
are feldspathic litharenites with mean values of Q41F25Lf34,
5.2. Results Qt41F23L36, and Qm33F23Lt44 (Figure 11a). This unit shows
[42] The Eocene La Paz Formation (4 samples) contains particularly high proportions of feldspar, dominantly plagio-
an average composition of Q73F12Lf15 (Figure 11a), defin- clase (Qm59P33K8), and volcanic fragments (Lm12Lv69Ls19).
ing a feldspathic litharenite to sublitharenite composition
[Folk, 1980]. Total lithic fragments (Lt)—defined as the 5.3. Interpretation
sum of metamorphic (Lm), volcanic (Lv), and sedimentary [48] Sandstone compositional data presented in ternary
(Ls) lithic fragments—show limited variations. La Paz diagrams (Figure 11) and a stratigraphic profile (Figure 12)
sandstones from the LSF hanging wall (Lm27Lv70Ls3) show compositional changes associated with Cenozoic shifts
display higher proportions of Lv to Lm relative to the in provenance.
footwall (Lm48Lv41Ls11). Compositional proportions accor- [49] 1. A marked increase in Ls in the uppermost Esmer-
ding to Dickinson and Suczek [1979] are Qt73F11L16 and aldas and Mugrosa Formations can be related to initial or
Qm60F12Lt28. enhanced sediment supply from the EC. This shift suggests
[43] The upper Eocene-lower Oligocene Esmeraldas For- that relative uplift of the EC was underway by late Eocene-
mation (4 samples) has an average composition of Q79F9Lf12 early Oligocene time. The low amount of Ls in the underly-
(with mean values of Qt77F9L14 and Qm68F9Lt23), which ing lower-middle Esmeraldas and La Paz suggests limited
straddles the subarkose, lithic arkose, and sublitharenite contributions from sedimentary sources during early middle
fields (Figure 11a). The proportion of Lm varies from the Eocene sedimentation, consistent with crystalline (igneous/
northern (Lm71Lv23Ls6) to southern footwall (Lm85Lv10Ls5), metamorphic) CC sources and/or sporadic supply from
but remains uniformly higher than the La Paz Formation. MMVB basement blocks such as the La Cira high [Moreno
[44] The Oligocene Mugrosa Formation (5 samples) has a et al., 2011]. Foraminiferal fragments in lower Miocene
sublitharenite average composition of Q88F4Lf8, with mean Colorado and upper Miocene Real samples indicate unroof-
values of Qt88F4L8 and Qm77F4Lt19 (Figure 11a). The amount ing of marine Cretaceous rocks of the EC, likely during LSF
of feldspar in this unit is very low (Qm95P2K3). Relative to motion.
underlying units, the Mugrosa also has a higher proportion of [50] 2. Increased compositional maturity in the Oligocene
polycrystalline quartz (Qp) and Ls (60% of Lt). However, Mugrosa and lower-middle Miocene Colorado Formations
an internal shift is expressed between the lower Mugrosa may represent sediment recycling from the EC, with attri-
(Lm6Lv15Ls79) and middle Mugrosa (Lm56Lv12Ls32). tion of unstable lithic fragments and feldspar. Low pro-
[45] The lower-middle Miocene Colorado Formation portions of Lv and Lm also may indicate a reduced supply
(10 samples) shows subarkose and sublitharenite compositions from the CC or MMVB basement highs concurrent with
with mean values of Q80F6Lf14, Qt80F6L14, and Qm76F6Lt22 increased EC contributions. Increased Ls agrees with facies
(Figure 11a). Whereas Ls is abundant (Lm2Lv31Ls67) in the relationships and paleocurrent data from the Nuevo Mundo
LSF footwall syncline, samples farther west show higher syncline 90 km to the north [Caballero, 2010; Caballero
Lm in lower levels (Lm64Lv8Ls28) with greater Ls in upper et al., 2010; Moreno et al., 2011].
levels (Lm2Lv34Ls64). Mudstones and siltstones dominate the [51] 3. A pronounced decrease in compositional maturity
Ls fraction, and there is a notable presence of foraminiferal from the lower-middle Miocene Colorado to upper Miocene
Ls in some lower Miocene Colorado and upper Miocene Real Group may indicate increased proximity of the EC
Real samples. source or recycling of the easternmost MMVB during LSF
[46] The upper Miocene Real Group (9 samples) consists motion. A coeval rise in sedimentation rates may further be
of litharenites and feldspathic litharenites, with mean values associated with enhanced Miocene-Pliocene shortening and
of Q55F8Lf36, Qt55F8L37, and Qm45F8Lt47 (Figure 11a). accommodation in the basin due to loading along its margins.

16 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

[52] 4. The high proportion of Lv and feldspar during the


past 5 Myr suggests a possible renewed volcaniclastic
supply from the CC that overprints the records of earlier EC
and MMVB provenance. This trend agrees with reports of
volcaniclastic sections correlative to this period in the Upper
Magdalena Valley basin [Wellman, 1970; Gómez et al.,
2003].
[53] Overall, Cenozoic units correspond to the recycled
orogen provenance suite according to compositional fields
defined by Dickinson and Suczek [1979] and Dickinson et al.
[1983]. However, the upper Miocene-Pliocene Mesa For-
mation straddles the dissected arc and recycled orogen fields
(Figure 11b and 11c), consistent with derivation from sedi-
mentary and crystalline sources associated with orogenesis.
A volcanic arc provenance attributable to the CC is most
pronounced for the Mesa and Real, but also characterizes the
La Paz, with limited proportions in the Esmeraldas, Mugrosa
and Colorado Formations. Although weathering and dia-
genesis affected sandstone compositions to some degree
by degrading unstable lithic fragments and feldspars (thus
enriching quartz content), we consider the major trends to
reflect changes in provenance due to the evolving composi-
tion of contributing source regions and adjustments in
drainage patterns induced by tectonic processes.

6. Kinematic Restorations
[54] The timing of deformation can be well constrained in
fold-thrust belts when pre-, syn- and post-tectonic strata are
preserved [e.g., Nemcok et al., 2005]. Where absent,

Figure 11. Ternary diagrams depicting the modal sandstone


compositions for 34 sandstone samples: (a) Q-F-Lf [Folk, Figure 12. Plot of major modal sandstone petrographic
1980], (b) Qt-F-L, and (c) Qm-F-Lt [Dickinson et al., 1983]. components according to stratigraphic level.

17 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

alternative approaches are required. Provenance studies may of the Opon anticline, a lower structural block (Figures 10h),
reveal changes in rock exposure generated by thrusting [e.g., shows deeper burial and initial cooling at 16–3 Ma.
Lonergan and Mange-Rajetzky, 1994; Qayyum et al., 2001; [59] A proposed model (Figure 13) explaining these
Nie et al., 2012]. Thermochronological techniques enable exhumation patterns, Ro temperature constraints, and sand-
estimates of the timing and duration of deformation [e.g., stone compositional results considers the structural evolution
Sinclair and Allen, 1992], with rates of fault displacement of the western EC and its link to footwall deformation in the
constrained by tracking exhumational cooling paths of MMVB. An incrementally restored cross section provides a
samples over a fault ramp [e.g., Huerta and Rodgers, 2006]. sequential eight-stage reconstruction (Figure 13).
In addition, Ro data help define pre-thrusting temperature [60] 1. Latest Cretaceous (65 Ma). This stage shows the
and aid reconstruction of thrust sheet geometry prior to pre-shortening geometry of a Mesozoic rift basin containing
faulting [e.g., Ohmori et al., 1997]. a >5 km thick Cretaceous succession. A Lower Cretaceous
[55] New structural, thermochronological, and provenance Ro sample in the east (sample 30) recorded a maximum
results of this study offer temporal constraints that make it burial depth of 6 km, comparable to a 5.4 km depth
possible to reconstruct the evolving configuration of differ- recorded by a significantly younger Upper Cretaceous hori-
ent structures and estimated rates of thrusting [e.g., zon in the west (sample 24). These results suggest a reduced
McQuarrie et al., 2005; Mora et al., 2008, 2010a, 2010b]. A sedimentary overburden in the east. We infer that the east-
sequentially restored cross section integrates the various data ernmost section represented an elevated block within the rift
sets and suggests a viable scenario that chronicles shortening system, consistent with regional interpretations for the Late
along the EC-MMVB boundary. Cretaceous basin configuration [Cooper et al., 1995;
Villamil, 1999, Gómez et al., 2003; 2005b]. Although direct
6.1. Methods evidence of the LSF as an ancestral normal fault is not
[56] From our structural restorations (Figures 5 and S5), observed, a Jurassic-cored hanging wall anticline farther
we selected a cross section lacking major oblique shortening. south (Figure S4b) is associated with an inversion structure.
Thermochronological, Ro, and sandstone petrographic data [61] 2. Paleocene–middle Eocene (65–45 Ma). Limited
were projected onto this section (Figure 9) and the restora- shortening of 2 km (2%) is compatible with Ro temperature
tion was modified to match constraints on paleotemperature estimates for the LSF hanging wall, which help constrain
and thrust sequence. Exhumation patterns were defined in maximum burial depths of 3–4 km (110–130 C) for the
both the footwall and hanging wall of the LSF. Ro data were Paleocene section in the western hanging wall, 4–5 km (160–
used to establish the pre-deformational geometry of the 180 C) for the Upper Cretaceous section in the central hanging
hanging wall and to calculate overburden thickness. Thermal wall and 5–6 km (180–220 C) for the Lower Cretaceous sec-
modeling was employed to help define the timing and tion in the eastern hanging wall. Thermochronological and Ro
sequence of thrusting. Sandstone compositions provided data indicate an eroded section about 3.5–4 km thick that
evidence about exhumation of structural blocks, in support originally capped the Paleocene in the frontal W edge of the
of thermochronological results. LSF hanging wall.
[62] 3. Middle Eocene-Oligocene (45–30 Ma). Partially
6.2. Results reset AFT results (sample 31; Figure 10c) indicate a possible
[57] Exhumation patterns inferred from thermal model- middle Eocene-Oligocene onset of cooling, likely controlled
ing of eight samples (Figure 10) indicate early exhumation by initial thrust displacement along the LSF, producing rock
(43–35 Ma) in the LSF hanging wall in the W forelimb of uplift in the frontal hanging wall. Additionally, a marked
the De Armas syncline (AS) (Figures 9 and 10c) and an AFT increase of sedimentary fragments in the Oligocene Mugrosa
PAZ situated in the Paleocene section. However, the E Formation suggests major sediment delivery from the EC.
backlimb of the AS (Figures 10b and 10d) shows later To the east, the Los Cobardes anticline may have been
exhumation (15–8 Ma) and deeper burial in which the growing by this time, as suggested by a non-reset ZFT sig-
Paleocene section was beneath the PAZ. Scarce, low-quality nature (sample 112), implying insufficient overburden to
AFT and ZFT samples for the Los Cobardes anticline (CA) induce thermal resetting.
(Figure 9) show that Jurassic rocks (sample 112) were not [63] 4. Oligocene–middle Miocene (30–15 Ma). Possible
beneath the ZFT closure temperature (250 C). These rela- LSF thrusting and splay faulting could control the differen-
tionships may suggest exhumation and initial growth of the tial exhumation patterns observed in the Opon syncline,
CA by Paleocene-Eocene time, thus precluding a thick where the W forelimb showed less overburden thickness
overburden. AFT results for the CA (sample 114) suggest during the Oligocene-late Miocene and a late onset of
progressive exhumation and growth of the anticline with the exhumation (8–3 Ma) for the upper Eocene Esmeraldas
Jurassic section cooling below the AFT closure temperature Formation (sample 18; Figure 10f). In contrast, the E back-
(120 C) at 6.7  1.6 Ma (Figure 9). limb recorded a greater Oligocene-lower Miocene overbur-
[58] LSF footwall exhumation patterns display slight dif- den and an early initiation of exhumation (18–10 Ma) for the
ferences on the flanks of the Opon syncline (OS) (Figure 9). Oligocene Mugrosa Formation (sample 17; Figure 10e).
Whereas 18–10 Ma cooling affected the Oligocene section Greater burial of the E flank of the Opon syncline could be
in the E backlimb adjacent to the LSF (Figure 10e), younger explained by propagation of a W-vergent LSF splay that
(11–3 Ma) cooling of the upper Eocene section in the fore- precluded deep burial of the leading edge; however, sedi-
limb indicates shallower burial (Figure 10f). However, the mentation may have exceeded the fault propagation rate
Miocene section in the OS shows modest cooling that did until late Miocene time. Earlier exhumation of the E flank
not reach the lower temperature of the PAZ (Figures 10g). could be associated with LSF displacement, suggesting that
The equivalent Miocene section farther west on the W flank

18 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

Figure 13. Kinematic restoration depicting Late Cretaceous to Pliocene deformation for the N-Open
section (Figures 4a and 5a), including the sequential activation of various faults (thick red lines) associated
with the La Salina fault (LSF). Results from thermochronological samples (orange dots) and vitrinite reflec-
tance (black dots) provide evidence to constrain deformation timing (see text for details). Approximate time
frames and cumulative shortening values for the different stages are listed in the lower left.

19 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

the tightly overturned E flank of the Opon syncline may 7.1. Geometry and Structural Style
have been influenced by footwall drag due to LSF thrusting. [70] Analyses of surface and subsurface geometries allow
[64] 5. Middle–late Miocene (15–10 Ma). An exhumation assessment of fold-thrust vergence patterns, degree of base-
difference across the LSF is inferred from a footwall thermal ment involvement, and the role of reactivation (inversion) of
model of the lower Colorado Formation, which displays Mesozoic normal faults during Cenozoic shortening. New
limited burial (60–70 C, presumably reaching the upper balanced cross sections supported by seismic evidence pro-
PAZ) (sample 20; Figure 10h), in contrast to the non-reset vide the basis for comparison with past studies of the EC-
(<50 C) hanging wall Colorado (sample 26; Figure 10g). MMVB boundary.
Growth of the E backlimb of the Opon syncline also [71] Along strike to the south of our study region, the
may have occurred by this time (cooling of sample 17; southern MMVB appears to share a similar geometry,
Figure 10e). In the LSF hanging wall, exhumation of the E shortening magnitude, and several-phase history of Cenozoic
backlimb of the De Armas syncline (samples 28 and 29; deformation and exhumation. Shortening was mostly
Figures 10b and 10d) may imply E-directed backthrusting accommodated by a basement-involved fault equivalent to
linked to displacement on a deeper W-directed thrust, with the LSF with a comparable displacement of 15–20 km
thrust-front propagation accommodated by tectonic wedging [Restrepo-Pace et al., 2004; Cortés et al., 2006; Sassi et al.,
in the Cretaceous section. 2007; Moretti et al., 2010]. Field mapping by Restrepo-Pace
[65] 6. Latest Miocene (10–6 Ma). Exhumation (sample et al. [2004] in the Rio Horta area 60 km south of our study
20; Figure 10h) could be associated with an E-directed back region revealed evidence of Paleocene shortening sealed
thrust observed in surface mapping and seismic profiles to by an early Eocene unconformity and a later phase of late
affect the Oligocene-Miocene section of the LSF footwall. Miocene tectonic wedging, attesting to at least two episodes
This back thrust accommodated westward advance of the of Cenozoic deformation. Regional kinematic models pro-
thrust front into the MMVB, causing recycling of the Mio- posed by Sassi et al. [2007] employed a passive-roof duplex
cene section on the W flank of the Opon anticline, with to transfer shortening from the EC to upper crustal levels
decreased compositional maturity of basin fill. of the MMVB, with an early W-directed structure predating
[66] 7. Latest Miocene-Pliocene (6–3 Ma). Continued MMVB Oligocene strata and a late Miocene structure
shortening induced exhumation of the Los Cobardes anti- involving large displacement along a major E-vergent
cline in the easternmost hanging wall at 6 Ma (AFT age for backthrust.
sample 114). Cooling also continued in the LSF hanging [72] These studies share notable similarities with our
wall (sample 14 at 12–6 Ma; Figure 10a) and in the MMVB interpretations, including comparable displacement along
(sample 18 at 8–3 Ma; Figure 10f). These thermal records a W-directed basement-involved thrust (LSF and its
represent exhumational cooling during rock uplift, possibly equivalents) with linked structures displaying a Cretaceous
associated with inversion and passive-roof duplexing of pre- detachment level. Discrepancies in shortening magnitude
Cretaceous to Neogene rocks along the W margin of the EC. could be related to model parameters such as decollément
[67] 8. Pliocene (3–0 Ma). Deformation culminated with depth, main thrust dip, and whether significant displace-
accelerated shortening (>2 km/Myr) and cumulative short- ment is accommodated by other faults. Possible explana-
ening of 27 km (25%). Plio-Quaternary shortening may tions for the apparent paradox of limited displacement
have been accommodated of out-of-sequence LSF motion along the leading (westernmost) thrusts of the MMVB
producing final structural relief (12 km). However, west- include (1) additional buried W- or E-directed thrusts [e.g.,
ward propagation along existing faults likely continued, Restrepo-Pace et al., 2004; Sassi et al., 2007], (2) a fault-
in addition to cutting of the Pliocene section by the small- propagation folding mechanism in which greater thrust dis-
displacement frontal thrust. placement is accommodated at depth [e.g., Suppe and
[68] To summarize, calculated shortening rates for the Medwedeff, 1990], or (3) a transpressional flower structure
kinematic restoration (Figure 13a) show a pronounced late with limited displacement on external splay faults [e.g.,
Miocene (7 Ma) increase in the rate of deformation to 1.5– Acosta et al., 2004, 2007]. Observations of shallowly
2 km/Myr (Figure 13b), in agreement with a reported late detached folds with steeply tilted Cenozoic strata overlying
Miocene-Pliocene climax of Andean deformation [Cooper gently dipping Cretaceous rocks and several E-directed
et al., 1995; Dengo and Covey, 1993; Cediel et al., 2003; backthrusts lead us to favor a passive-roof duplex to
Villamil, 1999; Mora et al., 2008]. Rates of Late Cretaceous accommodate additional EC shortening not transmitted to the
to middle Miocene shortening reveal slowest deformation MMVB [e.g., Banks and Warburton, 1986; Vann et al.,
(0.1 km/Myr) during the Oligocene-middle Miocene and 1986; Sassi et al., 2007]. By analogy, in the northernmost
moderate shortening (0.5 km/Myr) during the Eocene Andes along the Maracaibo mountain fronts, Duerto et al.
(Figure 13), which coincides with the proposed initiation of [2006] identified a triangular configuration of seismic
exhumation and rock uplift along the W margin of the EC. reflectors, shallow detachment levels within the Cretaceous-
Cenozoic section, and a basinward-dipping monocline with
7. Discussion essentially no surface faulting. We propose that a frontal
triangle zone associated with the main LSF thrust explains
[69] Integration of structural studies, basin analysis, and several structural elements: (1) a lower zone of imbricate
low-temperature thermochronology helps define the history basement thrusts; (2) an upper zone of basinward-dipping
of Cenozoic fold-thrust deformation and clastic sedimenta- backthrusts mainly localized in Upper Cretaceous levels; and
tion along the EC-MMVB transition. Below we consider (3) shortening accommodated by reactivation of normal
new evidence on the timing, style, and kinematics of defor- faults inherited from a Mesozoic rift. We further note that
mation in the northern Andes.

20 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

thin- and thick-skinned modes of deformation are not the which may attest to an effective topographic barrier pre-
products of separate tectonics events, but are the result of cluding sediment delivery from the CC.
kinematically linked ramp-flat and basement deformation
[e.g., Cooper et al., 1995; Mora et al., 2006].
8. Conclusions
7.2. Timing of Deformation and Exhumation
[80] The main conclusions of this study are summarized in
[73] Sediment provenance and fission track and (U-Th)/ the following six points.
He thermochronology help define the timing and pace of [81] 1. The eastern Middle Magdalena Valley basin
complex deformation in which vergence patterns are con- (MMVB) forms the proximal sector of an intermontane
sistent with tectonic wedging, backthrusting, and triangle hinterland basin largely controlled by W-directed thrusting
zone deformation. New results along the EC-MMVB tran- along the W edge of the Eastern Cordillera (EC). A proposed
sition are compared with previous studies to assess the onset, kinematic model suggests that shortening was accommodated
magnitude, and rates of deformation and exhumation. by W-directed basement thrusting and linked E-directed shal-
[74] Various data sets have been employed to evaluate the low backthrusting in a passive-roof duplex.
initiation of deformation and exhumation along the W margin [82] 2. Restoration of 4 cross sections along the EC-
of the EC (Figure S1). MMVB transition shows that >22–31 km of approximately
[75] 1. Growth strata indicate initial synorogenic sedi- E-W shortening was accommodated by reactivation of
mentation during the Eocene in well-studied parts of the inherited basement-involved faults. The Los Cobardes and
eastern MMVB to the south (Guaduas syncline) and north Portones anticlines in the western EC are cored by Jurassic
(Nuevo Mundo syncline) [Gómez et al., 2003, 2005a; synrift strata and separated by a relay zone, suggestive of
Moreno et al., 2011]. Selected regions show earlier defor- folding above inverted faults that once bounded a Jurassic-
mation, with Paleocene shortening suggested by map rela- Cretaceous extensional basin. Local non-plane strain
tionships in the Rio Horta area 60 km to the south responsible for area inconsistencies during the restoration
[Restrepo-Pace et al., 2004] and seismic relationships and could be related to minor oblique shortening.
fission track results for the northern Nuevo Mundo syncline [83] 3. Low-temperature thermochronology indicates three
[Parra et al., 2012]. main phases of Cenozoic cooling: (1) 45–30 Ma (middle
[76] 2. Low-temperature thermochronological results in Eocene–early Oligocene), associated with initial thrusting
the southern MMVB (Guaduas syncline) show initial along the La Salina fault (LSF); (2) 18–12 Ma (early middle
cooling between 50 and 35 Ma [Parra et al., 2009b], in Miocene), representing initiation of structures propagating
agreement with growth strata [Gómez et al., 2003]. Similar into the hinterland basin (MMVB); and (3) 12–3 Ma (late
data show younger cooling in the southern MMVB at 15– Miocene-Pliocene), a period of rapid LSF displacement and
5 Ma [Gómez et al., 2003]. In contrast, reset ZFT and out-of sequence deformation.
AFT ages of 24–19 Ma and 10–7 Ma, respectively, in the [84] 4. Vitrinite reflectance (Ro) results for 11 samples
western EC suggest relatively continuous exhumation support thermochronological data and provide information
along the main W-directed thrust fault equivalent to the about the initial basin configuration. Ro data suggest that the
LSF [Parra et al., 2009b]. However, several periods of Upper Cretaceous section in the LSF hanging wall achieved
exhumation are proposed to the north (Nuevo Mundo a burial depth of around 4–5 km during earliest synorogenic
syncline), consistent with growth strata evidence [Gómez sedimentation (Paleocene-Eocene), with diminished burial
et al., 2005a]. farther east in the central EC.
[77] 3. Detrital zircon U-Pb ages for the Nuevo Mundo [85] 5. Increased sedimentary lithic fragments suggest the
syncline show provenance signatures linked to EC exhu- development of sufficient EC topography to create a sedi-
mation between middle-late Eocene to Oligocene time [Nie ment source persisting since late Eocene-early Oligocene
et al., 2010]. time. Decreased compositional maturity during the Miocene-
[78] 4. Petrographic and paleoflow analyses suggest Pliocene may indicate westward growth (advance) of the EC
exhumation of the W flank of the EC by the middle-late source area during enhanced shortening and possible sedi-
Eocene [Caballero, 2010; Caballero et al., 2010; Moreno ment recycling of the easternmost MMVB.
et al., 2011; Saylor et al., 2011] in the Nuevo Mundo [86] 6. Structural reconstructions of the W flank of the EC
syncline and the late Eocene-Oligocene in the Guaduas show that the LSF was the principal structure responsible for
syncline [Gómez et al., 2003]. EC exhumation during the middle Eocene–early Oligocene
[79] These lines of evidence and new results from our (45–30 Ma). In contrast, exhumation since 18–12 Ma along
study suggest that the onset of exhumation for the W flank the EC-MMVB transition has been accommodated largely
of the EC occurred between the middle Eocene and early by tectonic wedging and E-directed backthrusting.
Oligocene. A significant exhumation episode in middle
Miocene-Pliocene time [Gómez et al., 2005a] inferred from [87] Acknowledgments. Funding was provided by the Ecopetrol-
ICP (Instituto Colombiano del Petróleo) project “Cronología de la deforma-
our thermal modeling may have caused significant rock ción en las Cuencas Subandinas” and the Jackson School of Geosciences at
uplift associated with EC basement provenance [Caballero, the University of Texas at Austin as part of a collaborative research agree-
2010; Caballero et al., 2010] as well as increases in sedi- ment. Additional funding was provided by a Geological Society of America
mentation rate and source-area proximity. These results are graduate student research grant. Seismic information, well data, and logisti-
cal support were provided by the aforementioned Ecopetrol-ICP project.
further consistent with the notable absence of magmatic-arc Apatite fission-track analyses for this Ecopetrol-ICP project were conducted
signatures in the middle Miocene-Pliocene section of the by Apatite to Zircon, Inc., and vitrinite reflectance data were generated by
Llanos foreland [Horton et al., 2010b; Bande et al., 2012], the ICP geochemistry lab. We thank Rocio Bernal, Henry Campos, Paul
Mann, Junsheng Nie, Mauricio Parra, and Joel Saylor for insightful

21 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

discussions. Constructive reviews by Isabelle Moretti, Fabio Speranza, and to tectonic setting, Geol. Soc. Am. Bull., 94, 222–235, doi:10.1130/0016-
Claudio Faccenna improved the quality of the manuscript. 7606(1983)94<222:PONAPS>2.0.CO;2.
Donelick, R. A., P. B. O’Sullivan, and R. A. Ketcham (2005), Apatite
fission-track analysis, Rev. Mineral. Geochem., 58, 49–94, doi:10.2138/
rmg.2005.58.3.
References Duerto, L., A. Escalona, and P. Mann (2006), Deep structure of the Merida
Acosta, J., L. Lonergan, and M. P. Coward (2004), Oblique transpression in Andes and Sierra de Perija mountain fronts, Maracaibo Basin, Venezuela,
the western thrust front of the Colombian Eastern Cordillera, J. South Am. Am. Assoc. Pet. Geol. Bull., 90, 505–528.
Earth Sci., 17, 181–194, doi:10.1016/j.jsames.2004.06.002. Duque-Caro, H. (1990), The Chocó Block in the northwestern corner of
Acosta, J., F. Velandia, J. Osorio, L. Lonergan, and H. Mora (2007), Strike- South America: Structural, tectonostratigraphic, and paleogeographic
slip deformation within the Colombian Andes, in Deformation of the implications, J. South Am. Earth Sci., 3, 71–84, doi:10.1016/0895-9811
Continental Crust: The Legacy of Mike Coward, edited by A. C. Ries, (90)90019-W.
R. W. H. Butler, and R. H. Graham, Geol. Soc. Spec. Publ., 272, 303–319. Etayo-Serna, F., et al. (1983), Mapa de terrenos geológicos de Colombia,
Bande, A., B. K. Horton, J. C. Ramírez, A. Mora, M. Parra, and D. F. Publ. Geol. Espec. Ingeominas, 14(1), 1–135.
Stockli (2012), Clastic deposition, provenance, and sequence of Andean Farley, K. A. (2000), Helium diffusion from apatite: General behavior as illus-
thrusting in the frontal Eastern Cordillera and Llanos foreland basin of trated by Durango fluorapatite, J. Geophys. Res., 105(B2), 2903–2914,
Colombia, Geol. Soc. Am. Bull., 124, 59–76, doi:10.1130/B30412.1. doi:10.1029/1999JB900348.
Banks, C. J., and J. Warburton (1986), ‘Passive-roof’ duplex geometry in the Farley, K. A., R. A. Wolf, and L. T. Silver (1996), The effects of long
frontal structures of the Kirthar and Sulaiman mountain belts, Pakistan, alpha-stopping distances on (U-Th)/He ages, Geochim. Cosmochim.
J. Struct. Geol., 8, 229–237, doi:10.1016/0191-8141(86)90045-3. Acta, 60, 4223–4229, doi:10.1016/S0016-7037(96)00193-7.
Bayona, G., M. Cortés, C. Jaramillo, and R. D. Llinás (2003), The Fusaga- Flowers, R., R. Ketcham, D. Shuster, and K. Farley (2009), Apatite (U–Th)/He
sugá succession: A record of the complex Latest Cretaceous–pre-Miocene thermochronometry using a radiation damage accumulation and annealing
deformation between the Magdalena Valley and Sabana de Bogotá areas, model, Geochim. Cosmochim. Acta, 73, 2347–2365, doi:10.1016/j.gca.
in VIII Simposio Bolivariano de Exploración Petrolera en las Cuencas 2009.01.015.
Subandinas, pp. 180–193, Asoc. Colombiana de Geol. y Geofis. del Folk, R. F. (1980), Petrology of Sedimentary Rocks, 184 pp., Hemphill,
Pet., Cartagena, Colombia. Austin, Tex.
Bayona, G., M. Cortés, C. Jaramillo, G. Ojeda, J. J. Aristizabal, and Garzanti, E., C. Doglioni, G. Vezzoli, and S. Ando (2007), Orogenic belts
A. Reyes-Harker (2008), An integrated analysis of an orogen–sedimentary and sediment provenance, J. Geol., 115, 315–334, doi:10.1086/512755.
basin pair: Latest Cretaceous–Cenozoic evolution of the linked Eastern Gómez, E., T. E. Jordan, R. W. Allmendiger, K. Hegarty, S. Nelly, and
Cordillera orogen and the Llanos foreland basin of Colombia, Geol. Soc. M. Heizler (2003), Controls on architecture of the Late Cretaceous to
Am. Bull., 120, 1171–1197, doi:10.1130/B26187.1. Cenozoic southern Middle Magdalena Valley Basin, Colombia, Geol.
Brandon, M. T., M. K. Roden-Tice, and J. I. Carver (1998), Late Cenozoic Soc. Am. Bull., 115, 131–147, doi:10.1130/0016-7606(2003)115<0131:
exhumation of the Cascadia accretionary wedge in the Olympic Moun- COAOTL>2.0.CO;2.
tains, northwest Washington State, Geol. Soc. Am. Bull., 110, 985– Gómez, E., T. E. Jordan, R. W. Allmendinger, K. Hegarty, and S. Kelley
1009, doi:10.1130/0016-7606(1998)110<0985:LCEOTC>2.3.CO;2. (2005a), Syntectonic Cenozoic sedimentation in the northern middle
Branquet, Y., A. Cheilletz, P. R. Cobbold, P. Baby, B. Laumonier, and Magdalena Valley Basin of Colombia and implications for exhumation
G. Giuliani (2002), Andean deformation and rift inversion, eastern edge of the Northern Andes, Geol. Soc. Am. Bull., 117, 547–569,
of Cordillera Oriental (Guateque-Medina area), Colombia, J. South Am. doi:10.1130/B25454.1.
Earth Sci., 15, 391–407, doi:10.1016/S0895-9811(02)00063-9. Gómez, E., T. E. Jordan, R. W. Allmendinger, and N. Cardozo (2005b),
Caballero, V. (2010), Evolucion tectono-sedimentaria del synclinal de Development of the Colombian foreland-basin system as a consequence
Nuevo Mundo, cuenca sedimentaria Valle Medio del Magdalena of diachronous exhumation of the northern Andes, Geol. Soc. Am. Bull.,
Colombia, durante el Oligoceno-Mioceno, MS thesis, 165 pp, Univ. 117, 1272–1292, doi:10.1130/B25456.1.
Ind. de Santander, Bucaramanga, Colombia. Hilley, G. E., and M. R. Strecker (2004), Steady state erosion of critical Cou-
Caballero, V., M. Parra, and A. Mora (2010), Levantamiento de la Cordil- lomb wedges with applications to Taiwan and the Himalaya, J. Geophys.
lera Oriental de Colombia durante el Eocene tardió–Oligoceno temprano: Res., 109, B01411, doi:10.1029/2002JB002284.
Proveniencia sedimentaria en el sinclinal de Nuevo Mundo, cuenca Valle Horton, B. K. (1999), Erosional control on the geometry and kinematics of
Medio del Magdalena, Bol. Geol. (Univ. Ind. Santander), 32, 45–77. thrust belt development in the central Andes, Tectonics, 18, 1292–1304,
Casero, P., J. F. Salel, and A. Rossato (1997), Multidisciplinary correla- doi:10.1029/1999TC900051.
tive evidences for polyphase geological evolution of the foothills of Horton, B. K. (2012), Cenozoic evolution of hinterland basins in the Andes
the Cordillera Oriental (Colombia), in VI Simposio Bolivariano de and Tibet, in Tectonics of Sedimentary Basins: Recent Advances, edited
Exploración Petrolera en las Cuencas Subandinas, pp. 100–118, Asoc. by C. Busby and A. Azor, pp. 427–444, Wiley-Blackwell, Oxford, U. K.
Colombiana de Geol. y Geofis. del Pet., Cartagena, Colombia. Horton, B. K., M. Parra, J. E. Saylor, J. Nie, A. Mora, V. Torres, D. F.
Cediel, F., R. P. Shaw, and C. Caceres (2003), Assembly of the Northern Stockli, and M. R. Strecker (2010a), Resolving uplift of the northern
Andean Block, in The Circum-Gulf of Mexico and the Caribbean: Hydro- Andes using detrital zircon age signatures, GSA Today, 20, 4–10,
carbon Habitats, Basin Formation, and Plate Tectonics, edited by doi:10.1130/GSATG76A.1.
C. Bartolini, R. T. Buffer, and J. Blickwede, AAPG Mem., 79, 815–848. Horton, B. K., J. E. Saylor, J. Nie, A. Mora, M. Parra, A. Reyes-Harker, and
Colletta, B., F. Hebrard, J. Letouzey, P. Werner, and J. L. Rudkiewicz D. F. Stockli (2010b), Linking sedimentation in the northern Andes to
(1990), Tectonic Style and crustal Structure of the Eastern Cordillera basement configuration, Mesozoic extension, and Cenozoic shortening:
(Colombia) from a Balanced Cross Section, in Petroleum and Tectonics Evidence from detrital zircon U-Pb ages, Eastern Cordillera, Colombia,
in Mobile Belts, edited by J. Letouzey, pp. 81–100, Technip, Paris. Geol. Soc. Am. Bull., 122, 1423–1442, doi:10.1130/B30118.1.
Cooper, M. A., et al. (1995), Basin development and tectonic history of Hossack, J. R. (1979), Use of balanced cross-sections in the calculation
the Eastern Cordillera, Llanos Basin and Middle Magdalena Valley, of orogenic contraction: A review, J. Geol. Soc., 136, 705–711,
Colombia, AAPG Bull., 19, 1421–1443. doi:10.1144/gsjgs.136.6.0705.
Corredor, F. (2003), Eastward extent of the late Eocene-early Oligocene Huerta, A., and D. Rodgers (2006), Constraining rates of thrusting and ero-
onset of deformation across the Northern Andes: Constraints from the sion: Insights from kinematic thermal modeling, Geology, 34, 541–544,
northern portion of the Eastern Cordillera fold belt, Colombia, J. South doi:10.1130/G22421.1.
Am. Earth Sci., 16, 445–457, doi:10.1016/j.jsames.2003.06.002. Ingersoll, R. V., T. F. Bullard, R. L. Ford, J. P. Grimm, J. D. Pickle, and
Cortés, M., B. Colleta, and J. Angelier (2006), Structure and tectonics of the S. W. Sares (1984), The effect of grain size on detrital modes; a test of
central segment of the Eastern Cordillera of Colombia, J. South Am. the Gazzi-Dickinson point-counting method, J. Sediment. Petrol., 54,
Earth Sci., 21, 437–465, doi:10.1016/j.jsames.2006.07.004. 103–116.
Dahlstrom, C. D. A. (1969), Balanced cross-sections, Can. J. Earth Sci., 6, Ingersoll, R. V., A. G. Kretchmer, and P. K. Valles (1993), The effect of
743–757, doi:10.1139/e69-069. sampling scale on actualistic sandstone petrofacies, Sedimentology, 40,
Dengo, C. A., and M. C. Covey (1993), Structure of the Eastern Cordillera 937–953, doi:10.1111/j.1365-3091.1993.tb01370.x.
of Colombia; Implications for trap styles and regional tectonics, AAPG Kammer, A., and J. Sánchez (2006), Early Jurassic rift structures associated
Bull., 77, 1315–1337. with the Soapaga and Boyacá faults of the Eastern Cordillera, Colombia:
Dickinson, W. R., and C. A. Suczek (1979), Plate tectonics and sandstone Sedimentological inferences and regional implications, J. South Am.
compositions, AAPG Bull., 63, 2164–2182. Earth Sci., 21, 412–422, doi:10.1016/j.jsames.2006.07.006.
Dickinson, W. R., L. S. Beard, G. R. Brakenridge, J. L. Erjavec, R. C. Kellogg, J. (1984), Cenozoic tectonic history of the Sierra de Perija,
Ferguson, K. F. Inman, R. A. Knepp, F. A. Lindberg, and P. T. Ryberg Venezuela-Colombia, and adjacent basins, in The Caribbean-South
(1983), Provenance of North American Phanerozoic sandstones in relation

22 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

American Plate Boundary and Regional Tectonics, edited by W. E. Nie, J., B. K. Horton, J. E. Saylor, A. Mora, M. Mange, C. N. Garzione,
Bonini, R. B. Hargraves, and R. Shagam, Mem. Geol. Soc. Am., 162, A. Basu, C. J. Moreno, V. Caballero, and M. Parra (2012), Integrated
239–261. provenance analysis of a convergent retroarc foreland system: U-Pb ages,
Ketcham, R. A. (2005), Forward and inverse modeling of low-temperature heavy minerals, Nd isotopes, and sandstone compositions of the Middle
thermochronometry data, in Low Temperature Thermochronology: Tech- Magdalena Valley basin, northern Andes, Colombia, Earth Sci. Rev.,
niques, Interpretations, and Applications, Rev. Mineral. Geochem., 110, 111–126, doi:10.1016/j.earscirev.2011.11.002.
vol. 58, edited by P. W. Reiners and T. A. Ehlers, pp. 275–314, Mineral. O’Sullivan, P. B., and W. K. Wallace (2002), Out-of sequence, basement-
Soc. of Am., Washington, D. C., doi:10.2138/rmg.2005.58.11. involved structures in the Sadlerochit Mountains region of the Arctic
Ketcham, R. A., A. Carter, R. A. Donelick, J. Barbarand, and A. J. Hurford National Wildlife Refuge, Alaska: Evidence and implications from fis-
(2007), Improved modeling of fission-track annealing in apatite, Am. sion-track thermochronology, Geol. Soc. Am. Bull., 114, 1356–1378,
Mineral., 92, 799–810, doi:10.2138/am.2007.2281. doi:10.1130/0016-7606(2002)114<1356:OOSBIS>2.0.CO;2.
Lonergan, L., and M. A. Mange-Rajetzky (1994), Evidence for Internal Ohmori, K., T. A. Tokuyama, H. Sakaguchi, A. Okamura, and A. Aihara
Zone unroofing from foreland basin sediments, Betic Cordillera, SE (1997), Paleothermal structure of the Shimanto accretionary prism, Shikoku,
Spain, J. Geol. Soc., 151, 515–529, doi:10.1144/gsjgs.151.3.0515. Japan: Role of an out-of-sequence thrust, Geology, 25, 327–330,
Mann, P., A. Escalona, and M. Castillo (2006), Regional geologic and tec- doi:10.1130/0091-7613(1997)025<0327:PSOTSA>2.3.CO;2.
tonic setting of the Maracaibo supergiant basin, western Venezuela, Pardo-Trujillo, A., C. A. Jaramillo, and F. E. Oboh-Ikuenobe (2003), Paleo-
AAPG Bull., 90, 445–478. gene Palynostratigraphy of the Eastern Middle Magdalena Valley,
McCourt, W. J., J. A. Aspden, and M. Brook (1984), New geological and Colombia, Palynology, 27, 155–178.
geochronological data from the Colombian Andes:Continental growth Parra, M., A. Mora, C. Jaramillo, M. R. Strecker, E. R. Sobel, L. Quiroz,
by multiple accretion, J. Geol. Soc., 141, 831–845, doi:10.1144/ M. Rueda, and V. Torres (2009a), Orogenic wedge advance in the
gsjgs.141.5.0831. northern Andes: Evidence from the Oligocene-Miocene sedimentary
McQuarrie, N., B. K. Horton, G. Zandt, S. Beck, and P. G. DeCelles (2005), record of the Medina Basin, Eastern Cordillera, Colombia, Geol. Soc.
Lithospheric evolution of the Andean fold-thrust belt, Bolivia, and the Am. Bull., 121, 780–800, doi:10.1130/B26257.1.
origin of the central Andean plateau, Tectonophysics, 399, 15–37, Parra, M., A. Mora, E. R. Sobel, M. R. Strecker, and R. González (2009b),
doi:10.1016/j.tecto.2004.12.013. Episodic orogenic-front migration in the northern Andes: Constraints
Milliken, K. L. (1988), Loss of provenance information through subsurface from low-temperature thermochronology in the Eastern Cordillera,
diagenesis in Plio-Pleistocene sandstones, northern Gulf of Mexico, Colombia, Tectonics, 28, TC4004, doi:10.1029/2008TC002423.
J. Sediment. Petrol., 58, 992–1002. Parra, M., A. Mora, C. Lopez, L. E. Rojas, and B. K. Horton (2012), Detect-
Montes, C., P. A. Restrepo-Pace, and R. D. Hatcher Jr. (2003), Three- ing earliest shortening and deformation advance in thrust-belt hinterlands:
dimensional structure and kinematics of the Piedras-Girardot fold belt: Example from the Colombian Andes, Geology, 40, 175–178, doi:10.1130/
Surface expression of transpressional deformation in the northern Andes, G32519.1.
in The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habi- Pindell, J., and K. Tabbutt (1995), Mesozoic–Cenozoic Andean paleogeog-
tats, Basin Formation and Plate Tectonics, edited by C. Bartolini, R. T. raphy and regional controls on hydrocarbon systems, in Petroleum Basins
Buffler, and J. Blickwede, AAPG Mem., 79, pp. 849–873. of South America, edited by A. J. Tankard, R. Suarez, and H. J. Welsink,
Montes, C., R. D. Hatcher Jr., and P. A. Restrepo-Pace (2005), Tectonic AAPG Mem., 62, 101–128.
reconstruction of the northern Andean blocks: Oblique convergence Qayyum, M., A. Niem, and R. Lawrence (2001), Detrital modes and prov-
and rotations derived from the kinematics of the Piedras-Girardot area, enance of the Paleogene Khojak Formation in Pakistan: Implications for
Colombia, Tectonophysics, 399, 221–250, doi:10.1016/j.tecto.2004. early Himalayan orogeny and unroofing, Geol. Soc. Am. Bull., 113,
12.024. 320–332, doi:10.1130/0016-7606(2001)113<0320:DMAPOT>2.0.CO;2.
Mora, A., M. Parra, M. R. Strecker, A. Kammer, C. Dimaté, and F. Rodriguez Reiners, P. W., T. L. Spell, S. Nicolescu, and K. A. Zanetti (2004),
(2006), Cenozoic contractional reactivation of Mesozoic extensional struc- Zircon (U-Th)/He thermochronometry: He diffusion and comparisons
tures in the Eastern Cordillera of Colombia, Tectonics, 25, TC2010, with 40Ar/39Ar dating, Geochim. Cosmochim. Acta, 68, 1857–1887,
doi:10.1029/2005TC001854. doi:10.1016/j.gca.2003.10.021.
Mora, A., M. Parra, M. R. Strecker, E. R. Sobel, H. Hooghiemstra, Restrepo-Moreno, S. A., D. A. Foster, D. F. Stockli, and L. N. Parra-Sanchez
V. Torres, and J. Vallejo-Jaramillo (2008), Climatic forcing of asymmetric (2009), Long-term erosion and exhumation of the “Altiplano Antio-
orogenic evolution in the Eastern Cordillera of Colombia, Geol. Soc. Am. queno”, northern Andes (Colombia) from apatite (U-Th)/He thermochro-
Bull., 120, 930–949, doi:10.1130/B26186.1. nology, Earth Planet. Sci. Lett., 278, 1–12, doi:10.1016/j.epsl.2008.
Mora, A., T. Gaona, J. Kley, D. Montoya, M. Parra, L. I. Quiroz, G. Reyes, 09.037.
and M. Strecker (2009), The role of inherited extensional fault segmenta- Restrepo-Pace, P. A., F. Colmenares, C. Higuera, and M. Mayorga (2004),
tion and linkage in contractional rogenesis: A reconstruction of Lower A fold and-thrust belt along the western flank of the Eastern Cordillera of
Cretaceous inverted rift basin in the Eastern Cordillera of Colombia, Colombia-Style, kinematics, and timing constraints derived from seismic
Basin Res., 21, 111–137, doi:10.1111/j.1365-2117.2008.00367.x. data and detailed surface mapping, in Thrust Tectonics and Hydrocarbon
Mora, A., B. K. Horton, A. Mesa, J. Rubiano, R. A. Ketcham, M. Parra, Systems, edited by K. R. McClay, AAPG Mem., 82, 598–613.
V. Blanco, D. Garcia, and D. F. Stockli (2010a), Migration of Cenozoic Roeder, D., and R. Chamberlain (1995), Eastern Cordillera of Colombia:
deformation in the Eastern Cordillera of Colombia interpreted from fission Jurassic-Neogene crustal evolution, in Petroleum Basins of South America,
track results and structural relationships: Implications for petroleum sys- edited by A. J. Tankard, R. Suarez, and H. J. Welsink, AAPG Mem., 62,
tems, AAPG Bull., 94, 1545–1580. 633–645.
Mora, A., M. Parra, M. R. Strecker, E. R. Sobel, G. Zeilinger, C. Jaramillo, Rolon, L. F. (2004), Structural geometry of the Jura-Cretaceous rift of the
S. Ferreira Da-Silva, and M. Blanco (2010b), The eastern foothills of the Middle Magdalena Valley Basin-Colombia, MS thesis, 63 pp, Univ. of
Eastern Cordillera of Colombia: An example of multiple factors control- W. Va, Morgantown.
ling structural styles and active tectonics, Geol. Soc. Am. Bull., 122, Sarmiento, L. F. (2001), Mesozoic rifting and Cenozoic basin inversion his-
1846–1864, doi:10.1130/B30033.1. tory of the Eastern Cordillera, Colombian Andes. Inferences from tec-
Moreno, C. J., B. K. Horton, V. Caballero, A. Mora, M. Parra, and J. Sierra tonic models, PhD thesis, 295 pp, Vrije Univ., Amsterdam.
(2011), Depositional and provenance record of the Paleogene transition Sarmiento-Rojas, L. F., J. D. Van Wess, and S. Cloetingh (2006), Mesozoic
from foreland to hinterland basin evolution during Andean orogenesis, transtensional basin history of the Eastern Cordillera,Colombian Andes:
northern Middle Magdalena Valley Basin, Colombia, J. South Am. Earth Inferences from tectonic models, J. South Am. Earth Sci., 21, 383–411,
Sci., 32, 246–263, doi:10.1016/j.jsames.2011.03.018. doi:10.1016/j.jsames.2006.07.003.
Moretti, I., G. R. Charry, M. M. Morales, and J. C. Mondragon (2010), Inte- Sassi, W., R. Graham, R. Gillcrist, M. Adams, and R. Gómez (2007), The
grated exploration workflow in the south Middle Magdalena Valley impact of deformation timing on the prospectivity of the Middle Magda-
(Colombia), J. South Am. Earth Sci., 29, 187–197, doi:10.1016/j. lena sub-thrust, Colombia, in Deformation of the Continental Crust: The
jsames.2009.08.011. Legacy of Mike Coward, edited by A. C. Ries, R. W. H. Butler, and R. H.
Nemcok, M., S. Schamel, and R. Gayer (2005), Thrustbelts: Structural Graham, Geol. Soc. Spec. Publ., 272, 473–498.
Architecture, Thermal Regimes and Petroleum Systems, 541 pp., Cam- Saylor, J. E., A. Mora, B. K. Horton, and J. Nie (2009), Controls on the iso-
bridge Univ. Press, Cambridge, U. K., doi:10.1017/CBO9780511584244. topic composition of surface water and precipitation in the northern
Nie, J., B. K. Horton, A. Mora, J. E. Saylor, T. B. Housh, J. Rubiano, Andes, Colombian Eastern Cordillera, Geochim. Cosmochim. Acta, 73,
and J. Naranjo (2010), Tracking evolution of Andean ranges bounding 6999–7018, doi:10.1016/j.gca.2009.08.030.
the Middle Magdalena Valley basin, Colombia, Geology, 38, 451–454, Saylor, J. E., B. K. Horton, J. Nie, J. Corredor, and A. Mora (2011), Eval-
doi:10.1130/G30775.1. uating foreland basin partitioning in the northern Andes using Cenozoic

23 of 24
TC3008 SÁNCHEZ ET AL.: ANDEAN FOLD-THRUST KINEMATICS, COLOMBIA TC3008

fill of the Floresta basin, Eastern Cordillera, Colombia, Basin Res., 23, Toro, J. (1990), The termination of the Bucaramanga Fault in the Cordillera
377–402, doi:10.1111/j.1365-2117.2010.00493.x. Oriental, Colombia, MS thesis, 60 pp, Univ. of Ariz., Tucson.
Saylor, J. E., D. F. Stockli, B. K. Horton, J. Nie, and A. Mora (2012), Dis- Toro, J., F. Roure, N. Bordas-Le Floch, S. Le Cornec-Lance, and W. Sassi
criminating rapid exhumation from syndepositional volcanism using (2004), Thermal and Kinematic Evolution of the Eastern Cordillera fold-
detrital zircon double dating: Implications for the tectonic history of the and-thrust-belt, Colombia, in Deformation, Fluid Flow and Reservoir
Eastern Cordillera, Colombia, Geol. Soc. Am. Bull., 124, 762–779, Appraisal in Foreland Fold-and-Thrust Belts, AAPG Hedberg Ser., vol. 1,
doi:10.1130/B30534.1. edited by F. Roure and R. Swennen, pp. 79–115, Am. Assoc. of Pet. Geol.,
Schamel, S. (1991), Middle and Upper Magdalena basins, Colombia, in Tulsa, Okla.
Active Margin Basins, edited by K. T. Biddle, AAPG Mem., 52, 283–302. Trenkamp, R., J. N. Kellogg, J. T. Freymueller, and H. P. Mora (2002),
Shagam, R., B. P. Kohn, P. O. Banks, L. E. Dasch, R. Vargas, G. I. Wide plate margin deformation, southern Central America and northwest-
Rodríguez, and N. Pimentel (1984), Tectonic implications of Cretaceous– ern South America, CASA GPS observations, J. South Am. Earth Sci.,
Pliocene fission track ages from rocks of the circum-Maracaibo Basin region 15, 157–171, doi:10.1016/S0895-9811(02)00018-4.
of western Venezuela and eastern Colombia, in The Caribbean-South Amer- Vann, I. R., R. H. Graham, and A. B. Hayward (1986), The structure
ican Plate Boundary and Regional Tectonics, edited by W. E. Bonini, R. B. of mountain fronts, J. Struct. Geol., 8, 215–227, doi:10.1016/0191-
Hargraves, and R. Shagam, Geol. Soc. Am. Mem., 162, 385–412. 8141(86)90044-1.
Sinclair, H. D., and P. A. Allen (1992), Vertical versus horizontal motions Villamil, T. (1999), Campanian-Miocene tectonostratigraphy, depocenter
in the Alpine orogenic wedge: Stratigraphic response in the foreland evolution and basin development of Colombia and western Venezuela,
basin, Basin Res., 4, 215–232, doi:10.1111/j.1365-2117.1992.tb00046.x. Palaeogeogr., Palaeoclimatol., Palaeoecol., 153, 239–275, doi:10.1016/
Spikings, R. A., W. Winkler, D. Seward, and R. Handler (2001), Along- S0031-0182(99)00075-9.
strike variations in the thermal and tectonic response of the continental Wellman, S. (1970), Stratigraphy and petrology of the nonmarine Honda
Ecuadorian Andes to the collision with heterogeneous oceanic crust, Earth group (Miocene), Upper Magdalena Valley, Colombia, Geol. Soc. Am.
Planet. Sci. Lett., 186, 57–73, doi:10.1016/S0012-821X(01)00225-4. Bull., 81, 2353–2374, doi:10.1130/0016-7606(1970)81[2353:SAPOTN]
Spikings, R. A., W. Winkler, R. A. Hughes, and R. Handler (2005), 2.0.CO;2.
Thermochronology of Allochthonous Terranes in Ecuador: Unraveling Zhang, E., and A. Davis (1993), Coalification patterns of the Pennsylvanian
the accretionary and post-accretionary history of the Northern Andes, coal measures in the Appalachian foreland basin, western and south-
Tectonophysics, 399, 195–220, doi:10.1016/j.tecto.2004.12.023. central Pennsylvania, Geol. Soc. Am. Bull., 105, 162–174, doi:10.1130/
Suppe, J. (1983), Geometry and kinematics of fault-bend folding, Am. 0016-7606(1993)105<0162:CPOTPC>2.3.CO;2.
J. Sci., 283, 684–721, doi:10.2475/ajs.283.7.684. Zhou, Y., M. A. Murphy, and A. Hamade (2006), Structural development
Suppe, J., and D. A. Medwedeff (1990), Geometry and kinematics of fault- of the Peregrina-Huizachal anticlinorium, Mexico, J. Struct. Geol., 28,
propagation folding, Eclogae Geol. Helv., 83, 409–454. 494–507, doi:10.1016/j.jsg.2005.11.005.
Sweeney, J. J., and A. K. Burnham (1990), Evaluation of a simple model
of vitrinite reflectance based on chemical kinetics, AAPG Bull., 74,
1559–1570.

24 of 24

You might also like