You are on page 1of 33

Downloaded from http://sp.lyellcollection.

org/ at Dalhousie University on November 13, 2013

Geological Society, London, Special Publications Online First

Inversion tectonics under increasing rates of


shortening and sedimentation: Cenozoic example from
the Eastern Cordillera of Colombia
Andrés Mora, Andrés Reyes-Harker, Guillermo Rodriguez, Eliseo
Tesón, Juan Carlos Ramirez-Arias, Mauricio Parra, Víctor
Caballero, José Pedro Mora, Isaid Quintero, Víctor Valencia,
Mauricio Ibañez, Brian K. Horton and Daniel F. Stockli

Geological Society, London, Special Publications, first published


March 8, 2013; doi 10.1144/SP377.6

Email alerting click here to receive free e-mail alerts when


service new articles cite this article
Permission click here to seek permission to re-use all or
request part of this article

Subscribe click here to subscribe to Geological Society,


London, Special Publications or the Lyell
Collection
How to cite click here for further information about Online
First and how to cite articles

Notes

© The Geological Society of London 2013


Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

Inversion tectonics under increasing rates of shortening


and sedimentation: Cenozoic example from the Eastern
Cordillera of Colombia
ANDRÉS MORA1*, ANDRÉS REYES-HARKER1, GUILLERMO RODRIGUEZ1,
ELISEO TESÓN1, JUAN CARLOS RAMIREZ-ARIAS1, MAURICIO PARRA1,
VÍCTOR CABALLERO1, JOSÉ PEDRO MORA1, ISAID QUINTERO1,
VÍCTOR VALENCIA2, MAURICIO IBAÑEZ2,
BRIAN K. HORTON3 & DANIEL F. STOCKLI4
1
Instituto Colombiano del Petróleo-Ecopetrol, km 7 Via a Piedecuesta,
Bucaramanga, Colombia
2
Department of Geosciences, University of Arizona, Tucson, AZ 85721, USA
3
Department of Geological Sciences, Jackson School of Geosciences,
University of Texas at Austin, Austin, TX 78712, USA
4
Department of Geology, University of Kansas, Lawrence, KS 66045, USA
*Corresponding author (e-mail: andres.mora@ecopetrol.com.co)

Abstract: The Northern Andes of Colombia is a key locality for understanding tectonic inversion
of symmetric rifts. A review of available data on structural geometry and deformation timing, and
new thermochronology and provenance data from selected localities, enable the construction of
balanced cross-sections and shortening budgets. During early deformation in the Palaeocene,
most shortening was focused in the western sector of the orogen, in the Central Cordillera and
the Magdallena Valley, although widely spaced and mild inversion occur in areas as far to the
east as the Llanos Basin. After a period of tectonic quiescence in the Middle Eocene, deformation
resumed across a former early Mesozoic graben in the Eastern Cordillera. Peak shortening rates and
out-of-sequence reactivation of the main inversion faults were in place in latest Miocene time,
during a phase of topographical growth. Our results indicate that coeval activation of basement
highs and adjacent slower-slip shortcuts appear to be characteristic of inverted symmetric
grabens. However, before reactivation and brittle faulting occur, strain hardening is required.
Deformation rates in the Eastern Cordillera correlate with the westwards velocity of the South
American Plate. A threshold convergence rate of approximately 2 cm year21 seems to be necessary
to activate shortening in the upper plate.

In convergent margins with ocean –continent plate certainly been observed in typical situations such
boundaries, most conventional models suggest that as in the Neogene evolution of the sub-Andean
orogenesis in the upper plate advances towards the zones of Argentina (Echavarria et al. 2003) and
foreland region, which in the Andes would be a sys- Bolivia (Uba et al. 2009). However, in other situ-
tematic eastwards propagation of deformation ations, prestrained provinces become an integral
(Butler 1987; DeCelles & Mitra 1995; DeCelles & part of evolving orogens and the deformation
Horton 2003). At the scale of thrust belts this has history is influenced by inhomogeneous basement
been called a forward-breaking sequence, which is configurations, rapidly varying facies and thick-
characterized by a principal thrust sheet kinemati- nesses of sedimentary packages, and generally long-
cally linked to smaller thrust faults with less displa- lasting inherited anisotropies and structures (Ring
cement closer to the foreland basin. This precise 1994; Wetzel et al. 2003; Paton & Underhill
situation should be strictly true only in cases 2004). A key question concerns how deformation
where the push clearly comes from the hinterland migrates during contractional orogenesis in those
and deforms a homogeneous (layer-cake) sedimen- settings. More precisely, how do previous exten-
tary succession of shales and sands or other mechan- sional histories (i.e. the presence of extensional
ical stratigraphy with contrasting rheologies that basins) influence the locus and distribution of later
systematically taper towards the foreland. This has contractional deformation?

From: Nemčok, M., Mora, A. & Cosgrove, J. W. (eds) Thick-Skin-Dominated Orogens: From Initial
Inversion to Full Accretion. Geological Society, London, Special Publications, 377,
http://dx.doi.org/10.1144/SP377.6, updated version # The Geological Society of London 2013.
Publishing disclaimer: www.geolsoc.org.uk/pub_ethics
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

In the Northern Andes, the Eastern Cordillera is affinity constitute the Central Cordillera. Finally,
the most external and youngest component of the the Eastern Cordillera, which is the topic of this
Colombian Andes, and is therefore a lateral equival- study, comprises the easternmost external chain of
ent of the sub-Andean zone of the central Andes of the Colombian Andes, which is separated from the
Argentina and Bolivia (Mora et al. 2010a). The Central Cordillera by the Magdalena Valley inter-
most evident deformation phases in this chain montane hinterland basin (Fig. 1). The area of the
occurred in the Neogene (Cooper et al. 1995; Villa- present-day Eastern Cordillera mostly coincides
mil 1999) but recent data show even earlier defor- with a Neocominan extensional domain, where the
mation events (Parra et al. 2012; Caballero et al. areas of most-positive relief roughly coincide with
2013, this volume, in press). Although the Eastern the most subsident area during the Early Cretaceous
Cordillera’s structural position is similar to the sub- (Mora et al. 2006; Sarmiento-Rojas et al. 2006).
Andean ranges of the Central Andes and it consti- Taboada et al. (2000) suggested that the Late
tutes the active deformation front of the Andes at Miocene indentation of the Panamá– Baudó arch
this latitude, its pre-existing configuration and style was responsible for the main phase of mountain
of deformation is very different. In particular, the building in the Eastern Cordillera. This hypothesis
relatively gentle deformation and moderate exhu- has been recently supported based on the correla-
mation in comparison to other analogue orogens, tion of oroclinal bending of the Eastern Cordillera
and wealth of new data acquired by Ecopetrol S.A with the shape of the Panamá –Baudó arch (Mann
in the framework of the project ‘Cronologı́a de la & Vargas-Jimenez 2011). However, recent data
deformación en las Cuencas Subandinas’, make suggest that the tectonic collision of the Panamá
the Eastern Cordillera an ideal pilot study region. isthmus actually began by the Late Oligocene
The region is well suited for analysing the tempo- (Farris et al. 2011), which also correlates with the
ral and spatial distribution of deformation in inver- most updated views and recent data on the moun-
sion orogens regarding the polarity of deformation tain building in the Eastern Cordillera. This sug-
migration, the rates of shortening and attendant gests, in stark contrast with the earlier studies
boundary, and the degree of coupling between con- (Cooper et al. 1995; Casero et al. 1997), that moun-
tractional orogenesis and former extensional defor- tain building in the Eastern Cordillera affected
mation. We therefore review the available data and virtually all of the former Neocomian graben by
present new data sets regarding the structural styles, Late Oligocene time (Mora et al. 2010b; Silva
deformation history and rates in order to propose et al. this volume, in press). A key question is
a model of deformation migration in inversion oro- whether this phase, defined as a stage of faster oro-
gens and its controlling factors. To understand what genic advance in the Eastern Cordillera (Parra et al.
controls the observed patterns, we propose diverse 2009b), is also related to the earlier age proposed
role for plate tectonics, the properties of the rocks and for Panama collision (Farris et al. 2011).
materials that were deformed or deposited during Based on the analysis of earthquake focal mech-
deformation, and the role of surface processes. anisms, Cortés & Angelier (2005) identified a rough
boundary for the influence of the Nazca Plate v. the
Caribbean Plate in transferring stress to the Eastern
Tectonic setting Cordillera. This boundary is located at about 48N.
In that sector, there is a fundamental change from a
The Northern Andes has a different configuration NNE– SSW to an east –west to NNW –SSE stress
compared with the Central Andes, largely because direction.
it is conditioned by several major plates interacting
with the orogenic belt (i.e. Caribbean, Nazca and
South American plates). However, these plates The Eastern Cordillera
have only been in their present configuration since
about 23 Ma, when the Farallon Plate was broken As previously suggested (Mora et al. 2010c), the
into the Nazca and Cocos plates, which caused a Eastern Cordillera constitutes a weak block
reorganization of the Pacific plates (Gutscher et al. between the stronger cratonic domains of the
1999). This reorganization has also been suggested Garzon Massif and the Llanos to the SE, as well
as a key factor influencing plateau building in the as the Maracaibo block to the NE (Fig. 1). This
Central Andes (Allmendinger et al. 1997). strength contrast is reflected in the fact that these
The Northern Andes comprise three different regions underwent little deformation during the
branches north of 28N. Accreted and uplifted main Cenozoic orogenesis, as revealed in seismic
oceanic material constitutes the Western Cordillera, profiles of the Maracaibo (Escalona & Mann
which is the westernmost branch deformed since the 2006) and Llanos (Delgado et al. 2012) blocks, as
Late Cretaceous –Palaeocene (Fig. 1). Uplifted well as by surface geology in the Garzón Massif.
igneous and metamorphic rocks of continental In contrast, the Eastern Cordillera appears to
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 1. Topographical map (GTOPO30) of NW South America showing major tectonomorphic elements.

coincide with a long lasting mobile belt bounding The Eastern Cordillera is a bivergent orogen that
the adjacent cratonic belts (Cooper et al. 1995; Mora coincides with a Neocomian depocentre which ori-
et al. 2006, 2010c; Sarmiento-Rojas et al. 2006). ginated during Early Cretaceous extension. Casero
Mora et al. (2010c) further refined the indenter et al. (1997) suggested that the Early Cretaceous
model of Taboada et al. (2000) by mapping the rift essentially coincides with the modern areas
main boundary faults of the weak block of the with the most positive relief in the Eastern Cordil-
Eastern Cordillera, which are the Tesalia –Algeciras lera. The rift prompted the fundamental bivergent
(Velandia et al. 2005; Mora et al. 2010c) and Bucar- configuration of the Eastern Cordillera, with two
amanga faults (Fig. 2). These two faults accommo- foothills-style marginal thrust belts, the Magdalena
date dextral and sinistral strike-slip motion, Valley belt to the west and the eastern foothills
respectively. Tesón et al. (this volume, in press) belt to the east, with dominantly west- and east-
presents multiple data reinforcing the idea of differ- vergent faults, respectively (Fig. 2a). As shown by
ent stress domains in the Eastern Cordillera, as Mora et al. (2006) and Tesón et al. (this volume,
suggested by Cortés et al. (2006) and Mora et al. in press), the typical association of both the west-
(2010c). In the updated indenter model (Mora ern and eastern foothill belts is that of a basement
et al. 2010c), the different stresses that actually high uplifted by reverse faults, usually inverted
cause the oroclinal bending in the Cocuy region normal faults, adjacent to thin-skinned belts coin-
are formed by the escape of the weaker Eastern ciding with Tertiary depocentres. At both the west-
Cordillera between the strong Maracaibo and ern and eastern mountain fronts, the thin-skinned
southern Llanos cratonic blocks (Figs 1 & 2b). belts are arranged in an en echelon pattern. Also
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 2. (a) Geological and (SRTM 30; b) topographical maps of the Eastern Cordillera, and the Middle Magdalena
and Llanos basins of Colombia. A genetic classification of the main faults is provided based on Mora et al.
(2006, 2010c), Caballero et al. (2013), Jimenez et al. (this volume, in press) and Tesón et al. (this volume, in press).

significant is the presence of a central, topographi- data sets that include growth strata in key locations,
cally elevated structural depression, the Sabana de new apatite fission-track thermochronology, as well
Bogotá (Bogotá Plateau), which is bounded by as cross-cutting field relationships. Based on this
two basement highs, the Farallones Anticline and line of evidence, we compiled maps of the timing
the Villeta Anticlinorium (Fig. 2). Basement expo- and style of the structures. Then we used data
sures in the weaker domain of the Eastern Cordillera bearing on the chronology of deformation to kine-
are restricted to the Floresta and Quetame basement matically restore the cross-sections presented in
uplifts (Fig. 2a). However, analogue basement- this Special Publication (Tesón et al. this volume,
involved structures, bounded by reverse (inverted) in press). This restoration was based on the retro-
faults but lacking basement exposures, include the deformation of growth strata, where observed,
Los Cobardes, Peñon and Arcabuco anticlines and on sediment provenance, syn-tectonic facies,
(Fig. 2a). The stratigraphy of the Eastern Cordillera palaeocurrents and mostly thermochronological
is summarized in Figure 3. constraints previously discussed in other studies
(Table 1), as well as the new one presented here
Methods (Tables 2 & 3). The principal goal was to try to
build high-resolution, calibrated thermal histories
We constructed the present study based on a using the software HeFTy (Ketcham 2005), and
global view of the available data on the structural incorporating vitrinite reflectance (Ro) data, apatite
geometry and timing of deformation from other and zircon fission-track data (AFT and ZFT,
chapters of this Special Publication (e.g. Caballero respectively) and (U –Th)/He data from this study,
et al. 2013, this volume, in press; Jimenez et al. published works and other chapters of this Special
this volume, in press; Moreno et al. this volume, Publication, taking the most representative results
in press; Silva et al. this volume, in press; Tesón from the different structural domains. Later, the
et al. this volume, in press), as well as on additional different faults were restored to eight different
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 3. Chronoestratigraphic framework along a transect across the Eastern Cordillera (shown in the inset map).
Modified from Gómez et al. (2003, 2005), Mora et al. (2006, 2010b) and Parra et al. (2009a, b, 2010).

steps during the Cenozoic based on the fact that the steps, as combined with the data from other kin-
interpreted geometries and fault trajectories should ematic restorations in this Special Publication,
be retrodeformed by an amount that satisfies the were used to calculate incremental shortening rates
one-dimensional HeFTy thermal histories and for each section. Finally, we compare our results
other geological constraints derived mostly from against different controlling factors for the timing,
growth strata and provenance data. These different style and distribution of deformation to propose a
Table 1. Apatite fission-track data used in this study

Sample Longitude Latitude Elevation Unit AFT age # MTL (mm) # Source
(8W) (8N) (m) (see ages in Fig. 3) (Ma) Gr +1s error† length

Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013


+1s error*

Figure 8
GC1093-64 73821′ 36.12′′ 4805′ 31.2′′ 23353 Chipaque Formation 31.7 + 3.1 20 12.02 + 0.22 100 This work
GC1093-67 73821′ 36.12′′ 4805′ 31.2′′ 23505 Une Formation 36.8 + 7.1 6 11.92 + 0.35 100 This work
Section Figure 14
606 –7 72812′ 09.92′′ 7800′ 19.06′′ 550 Carbonera (C8 –C6) 3.8 + 0.6 38 13.38 + 0.94 16 This work
Formation
′ ′′ ′ ′′
606 –11 72813 18.23 7808 41.53 621 Mirador Formation 5.0 + 0.9 28 13.64 + 0.59 7 This work
996 –04 73801′ 15.70′′ 7817′ 05.11′′ 1405 Tablazo Formation 10.8 + 3.5 9 12.84 + 0.91 7 This work
996 –20 73809′ 28.57′′ 7803′ 41.67′′ 753 Bucaramanga terraces 36.8 + 7.1 20 12.78 + 0.35 74 This work
996 –30 73806′ 58.27′′ 7809′ 16.57′′ 801 Palermo Formation 48.2 + 4.6 12 13.71 + 0.25 9 This work
1032 –15 72809′ 10.49′′ 7802′ 02.86′′ 381 Guayabo Group 14.1 + 1.3 34 12.89 + 0.26 39 This work

A. MORA ET AL.
1032 –19 73818′ 59.86′′ 7806′ 10.86′′ 1869 Une Formation 3.0 + 0.5 20 12.41 + 0.00 1 This work
1032 –24 72831′ 07.72′′ 7816′ 35.79′′ 1506 Floresta Formation 4.9 + 0.6 40 13.65 + 0.24 62 This work
Figure 15
RG-06 73820′ 54.42′′ 5836′ 39.00′′ 3121 Socha Formation 46.8 + 3.3 18 13.50 + 0.82 11 Parra et al. (2009a, b)
RON-1002 73809′ 46.96′′ 5821′ 56.40′′ 2055 Upper Socha Formation 42.0 + 4.2 15 13.32 + 0.22 34 Silva et al. (this volume, in press)
US-1004 74808′ 30.30′′ 4831′ 04.88′′ 2776 Socha Formation 52.8 + 2.5 37 14.07 + 1.12 200 Silva et al. (this volume, in press)
AFT-250710 – 12 72820′ 45.99′′ 6832′ 01.54′′ 4499 Mirador Formation 6.7 + 1.3 14 13.83 + 0.44 14 Silva et al. (this volume, in press)
MM-005 72831′ 28.38′′ 5828′ 52.46′′ 865 Guadalupe Formation 11.2 + 1.9 12 11.83 + 0.86 10 This work
CS-3 73842′ 55.44′′ 4811′ 53.62′′ 995 Guatiquia Formation 3.1 + 0.5 40 11.2 + 1.9 This work
Section Figure 17
RG-04 73805′ 17.59′′ 5855′ 30.02′′ 3619 Montebel Formation 19.8 + 2.1 38 13.61 + 0.22 53 Parra et al. (2009a, b)
AM-12 72853′ 42.99′′ 5849′ 26.74′′ 2550 Pre-Devonian basement 16.0 + 3.0 31 13.89 + 0.21 55 Parra et al. (2009a, b)
AM-09 72851′ 36.41′′ 5849′ 41.96′′ 2435 Une Formation 25.7 + 2.1 30 12.91 + 0.21 75 Parra et al. (2009a, b)
MP-27 72831′ 20.31′′ 5830′ 22.28′′ 839 Las Juntas Formation 11.7 + 2.5 24 14.03 + 0.58 8 Mora et al. (2010b)
MP-40 72830′ 04.01′′ 5829′ 3.1′′ 665 Une Formation 8.3 + 1.2 24 12.66 + 0.88 10 Mora et al. (2010b)
AM-02 72828′ 26.72′′ 5828′ 1.92′′ 639 Los Cuervos Formation 42.2 + 3.7 38 11.65 + 0.14 201 Mora et al. (2010b)
MP-82 72815′ 54.68′′ 5840′ 07.81′′ 582 Carbonera Formation 25.9 + 2.3 25 10.89 + 0.32 53 Mora et al. (2010b)
MP-85 72815′ 37.91′′ 5840′ 11.50′′ 564 Los Cuervos Formation 10.9 + 2.1 40 11.78 + 0.31 55 Mora et al. (2010b)
MP-72 72813′ 41.88′′ 5840′ 02.51′′ 486 Guayabo Group 22.8 + 3.4 23 14.80 + 1.14 45 Mora et al. (2010b)
MP-45 72807′ 42.42′′ 5837′ 31.96′′ 317 Guayabo Group 30.9 + 3.1 24 12.40 + 0.22 48 Mora et al. (2010b)

*
Pooled (central) age reported for ages that pass (fail) the x2 test.

Mean track length (MTL) of confined fission tracks measured.
Table 2. New apatite fission-track data acquired using the external detector method (EDM)

Sample Latitude Long Elevation Unit Stratigraphic # U Rho-S Rho-I Rho-D P(x2)‡ AFT age Cl MTL +1s SD #

Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013


(8N) itude (8W) (m) age (Ma) Gr (ppm) (NS)* (NI)* (ND)† (Ma) (%wt) error (mm)} (mm) length
+ 1s error§

1093 – 73821′ 36.12′′ 4805′ 31.2′′ 23353 Chipaque 94 –85 20 62 8.61 (267) 64.18 (1991) 11.86 (1860) ,1 31.7 + 3.1 0.00 –0.40 12.02 + 0.22 2.23 100
64 Formation
1093 – 73821′ 36.12′′ 4805′ 31.2′′ 23505 Une 106 –94 6 136 22.50 (126) 141.4 (792) 11.84 (1860) ,1 36.8 + 7.1 0.00 –0.30 11.92 + 0.35 2.22 100
67 Formation
MM-005 72831′ 28.38′′ 5828′ 52.46′′ 865 Guadalupe 82 –71 13 32 1.02 (37) 23.14 (836) 12.84 (2020) 92 11.2 + 1.9 0.01 –0.19 11.83 + 0.86 2.71 10
Group

INVERSION TECTONICS: COLOMBIAN ANDES


*
Rho-S and Rho-I are the spontaneous and induced track density measured, respectively (105 tracks cm22). NS and NI are the number of spontaneous and induced tracks counted for estimating Rho-S and
Rho-I, respectively.

RhoD is the induced track density measured in the external mica detector attached to CN2 dosimetry glass (105 tracks cm22). ND is the number of induced tracks counted in the mica for estimating RhoD.
‡ 2
(x ) (%) is the chi-square probability (Galbraith 1981; Green 1981). Values greater than 5% are considered to pass this test and represent a single population of ages.
§
Pooled (central) age reported for ages that pass (fail) the x2 test.
}
Mean track length of confined fission tracks measured.
Data produced and counted by Geotrack using a Zeta value of 392.9 + 7.4 for CN5 glass.

Table 3. New apatite fission-track data acquiered using laser abaltion-inductively coupled plasma-mass spectrometry (LA-ICP-MS)

Sample Latitude Longitude Eleva- Unit Strati- # Gr NS* S(PV)† 1sS(PW) jMS‡ 1s 43
Ca§ U}
238
P(x 2) AFT age Dpar SD #Dpar MTL + SD #
(8W) (8W) tion graphic (cm2) (E-07 cm2) jMS (Ma) + (mm) (mm) 1s error (mm) length
(m) age (Ma) 1s error (mm)**

CS-3 73842′ 4811′ 995 Guatiquia 360 –300 40 40 9.27E-05 2.99E-06 16.42 0.34 3.42E 2 02 2.44E 2 03 16 3.1 + 0.5 1.62 0.22 40 13.18 + 2.44 21
55.44′′ 53.62′′ Formation 0.53

*
Number of spontaneous fission tracks counted over area W.

Sum of Ri* Wi for all grains evaluated; Ri is (238U/43Ca) for apatite grain i; Wi is the area over which NS and Ri are evaluated.

z-calibration factor based on LA-ICP-MS of fission-track age standards.
§
Background-corrected 43Ca (dimensionless).
}
Background-corrected 238U (dimensionless).
 2
x (%) is the chi-square probability. Values greater than 5% are considered to pass this test and represent a single population of ages.
**
Mean track length of confined fission tracks measured.
# Gr, Number of apatite grains counted.
Dpar, Diameter parallel to the etch pit.
# length, number of lengths measured.
+1s, Error range.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 4. Pre-Cretaceous outcrops in the Central and Eastern cordilleras. Precambrian and Palaeozoic basement include
medium- and low-grade metamorphic rocks exposed in the Central Cordillera and the eastern flank of the Eastern
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

model of deformation that we suggest can be gener- west of the Boyacá Fault (Silva et al. this volume,
alized to the initial stages of many inversion orogens in press) (Fig. 5a, b). From this evidence, it is impor-
but more precisely to inversion orogens where short- tant to note that the areas of Palaeocene exhumation
ening and sedimentation rates increase with time. and important erosion coincide with the areas of
Permo-Triassic metamorphic basement and oldest
(Jurassic) syn-rift deposition (compare Figs 4 &
Results 5). However, important evidence by Bayona et al.
Basement framework (2013) from provenance, heavy minerals, reworked
pollen and glauconite, as well as stratigraphic thick-
As shown by Tesón et al. (this volume, in press), ness changes, indicate that to some extent there were
the Eastern Cordillera can be subdivided according probably minor intrabasinal highs in the foredeep
to its main basement domains (Fig. 4), which con- areas of the eastern portions of the former graben
ditioned later deformation patterns. The most signifi- basin undergoing inversion. It is our point of view
cant contrast is the fact that the Eastern Cordillera that those local highs were not detected by bed-
forms a part of two fundamental provinces. One pro- rock or detrital thermochronology because there
vince is located to the west, where metamorphic was limited structural and topographical relief
basement of Permo-Triassic age (Fig. 4c) underlies such that the uplift of these structural highs east of
Jurassic early syn-rift rock units (Fig. 4d). This the Boyacá Fault did not produce significant
province is located west of the Boyacá and Bucara- cooling effects in the rock units.
manga faults, and includes the documented sub- A new perspective from our paper and even more
surface relationships of the Middle Magdalena intriguing is the finding that most probably there
Valley Basin (Ibañez-Mejia et al. 2011). In contrast, was Palaeocene deformation in the flat-lying
the basement province to the east includes areas Llanos Basin (Fig. 6). This is interpreted first based
where the metamorphic basement is of early Palaeo- on the almost total absence of a Palaeocene strati-
zoic age (Horton et al. 2010b) (Fig. 4b) and under- graphic record (which pinches out very close to the
lies upper Palaeozoic sedimentary rocks (Fig. 4c). present-day deformation front) in the subsurface of
In those areas, dated Jurassic syn-rift rocks are the southern portions of the Llanos foreland basin.
absent and the earliest syn-rift deposits are mostly This is first documented in the public domain here
of Berriasian age (Ingeominas 2005). Such relation- in this work, which contrasts with the eastwards-
ships are well expressed in the Farallones Anticline advancing pinch-out of the Eocene –Neogene units
of the Quetame Massif, but also in the Santander (depositional zones in Figs 5c, d & 6). Second, pro-
and Floresta massifs. It is worth noting that with venance data indicate sediment derivation from
the notable exception of the Santander Massif (Fig. immature unstable sources that cannot be located
4a), unambiguous Precambrian outcrops are absent further than 150 km (Johnsson et al. 1991). Third,
in this domain, where the metamorphic rocks are exposed unconformities in the area of the Guamal
constituted by metasedimentary units. These obser- River show the Eocene Mirador Formation rest-
vations describe contrasting thermal and deforma- ing upon Late Cretaceous units with no Palaeocene
tion histories in Upper Palaeozoic and early Meso- facies preserved (Fig. 7). Fourth, our new thermo-
zoic rocks cropping out across the Boyacá and chronology in this study (Fig. 8), which allows
Bucaramanga faults. In the Permo-Triassic, burial Palaeocene cooling of exposed basement highs like
to metamorphic conditions in the west occurred the La Macarena High (Fig. 2) and buried highs
coevally to the accumulation of platformal sequences like the Melon el Viento. All of these lines of evi-
to the east. Subsequent Jurassic rift-related accu- dence allow us to suggest that most of the Pre-Oligo-
mulation only occurred in the western province. cene deformation visible on the seismic sections in
the Llanos Basin (see Fig. 6) is actually of Palaeo-
Earliest deformation cene age, similar to the Middle Magdalena Basin
(Caballero et al. 2013, this volume, in press).
Data from Parra et al. (2012) and Caballero et al.
(2013, this volume, in press) show a clear exhuma- Tectonic quiescence
tion event throughout the Palaeocene that probably
started as early as the latest Cretaceous. This event We suggest here that evidence in the subsurface of
has been documented in the subsurface of the the Middle Magdalena Valley area indicates that
Middle Magdalena Basin and also in outcrop areas Palaeocene deformation continued into the Early

Fig. 4. (Continued) Cordillera. Late Palaeozoic exposures show a marked contrast across the Bucaramanga and Boyacá
faults, with medium-grade metamorphic rocks to the west and platformal sedimentary rocks to the east. Jurassic
rift-related redbeds occur only west of these structures. See the text for further explanation and references.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 5. Palaeocene –Eocene palaeogeography of the study area based on surface and subsurface cross-cutting
stratigraphic relationships, and thermochronometric and provenance data. See Caballero et al. (2013) and Silva et al.
(this volume, in press), and references therein, for supporting information.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 6. Seismic reflection line from the Llanos Basin showing a basement high beneath undeformed Cenozoic strata.
Folded Cretaceous units uncomfomably lying beneath Oligocene strata constrain deformation to have occurred
prior to the Oligocene but during the Cenozoic Andean orogeny.

Eocene (Fig. 5c). However, we suggest that the Mid- significantly low sediment accumulation rates in
Eocene was a phase in which tectonic quiescence the Mid-Eocene Upper Mirador Formation in the
dominated the study area. We use three arguments eastern foothills (which represents the marine
to support this interpretation. First, the presence of fine-grained interval of the Mirador Formation)
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 7. Geological map of the Rio Guamal region showing cross-cutting relationships of Early Cenozoic deformation
in the Boa Anticline. The arrow indicates the locality where early Eocene strata of the Mirador Formation lies
unconformably on top of the Upper Cretaceous strata of the Guadalupe Group.

between two periods of higher sediment accumu- Eastern Cordillera in the entire Cenozoic record of
lation (Fig. 9). Second, westwards retrogradation the Nuevo Mundo Syncline, which we consider
of the Mid-Eocene provenance divide documented combined with the previous evidence as an indicator
by Silva et al. (this volume, in press) in the axial of slow tectonics or quiescence.
zone of the Eastern Cordillera (compare Fig. 5c,
d). In this case, the provenance divide returned to
a position closer to the Early Palaeocene deforma- Widespread exhumation, strain hardening
tion front, in the present-day hanging wall of the and low deformation rates
Boyacá Fault. We interpret this as the effect of an
increase in the ratio of accommodation space to With lag-time analysis of detrital zircons (Fig. 10),
sediment supply because of a lower sediment sup- Saylor et al. (2012) proposed a further advance of
ply from a stable area. Finally, in the Magdalena the deformation front in the Late Eocene after tec-
Valley, the fine-grained lower interval of the tonic quiescence, consistent with previous results
Middle Eocene Esmeraldas Formation (see Cabal- (Parra et al. 2009b; Mora et al. 2010b). At this time,
lero et al. 2013, this volume, in press) was also motion along the Pesca– Soapaga Fault System
deposited under lower sedimentation rates com- actively exhumed rocks in its hanging wall. Silva
pared to the previous and subsequent time intervals. et al. (this volume, in press) confirm that this
The detrital zircon provenance record of this unit domain of Late Eocene exhumation extends along
also documents an eastwards retreat of the prove- the entire hanging wall of the Machetá Fault south
nance divide between detrital zircons coming from of the Pesca –Soapaga Fault System (Fig. 5d).
the Central Cordillera and detrital zircons coming Several recent studies demonstrate that exhuma-
from the Eastern Cordillera. In fact, we find in this tion was ongoing by the Oligocene in the domain
interval the closest provenance divide to the of the entire ancestral graben (Parra et al. 2009b;
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

and had lower amplitude and shorter wavelength


involving a shallow (c. 5 km depth) detachment
(Kammer 1997; Kammer & Mora 1999; Mora &
Kammer 1999). In this study, we complement
these observations by showing growth strata to
document the behaviour in the present foothill prov-
inces (i.e. the Magdalena and eastern foothills in
Fig. 12). What we observe in virtually all of the mar-
ginal Tertiary synclines of both the western and
eastern foothills of the Eastern Cordillera is
Oligocene growth strata continuous with ongoing
deformation or minor unconformities where sedi-
mentation stopped briefly but quickly resumed in
the presence of active deformation (Gómez et al.
2003; Bayona et al. 2008; Parra et al. 2010). In
fact, Suppe et al. (1991) documented similar
growth unconformities that represent a near-zero
time hiatus.
The Oligocene was, therefore, a period of gener-
alized deformation throughout the graben domain,
with second-order contractional folds forming
across the entire graben basin (Tesón et al. this
volume, in press). The most significant new obser-
vation from our data is that, while the graben
domain was undergoing thrust-induced exhumation,
the foothill depocentres experienced almost zero
exhumation but recorded very high sedimentation
rates while deformation was ongoing as well.
Evidence in other chapters in this Special Publi-
cation (see Caballero et al. 2013; Jiménez et al. this
volume, in press; Moreno et al. this volume, in
Fig. 8. Thermal models obtained using HeFTy software press) and previous publications (Parra et al.
(Ketcham 2005) for Cretaceous susurface sedimentary 2009a, b; Horton et al. 2010a; Mora et al. 2010b;
rocks retrieved from the Merey-1 well. The dark-grey Bande et al. 2012; Ramı́rez-Arias et al. 2012)
area encloses time– temperature paths with a good-fit to show that the important basement high (mainly the
the measured AFT data. The light grey area encloses paths Farallones and Pisba anticlines in the eastern foot-
with only an acceptable fit. Black boxes correspond to
hills, and the Cobardes and Villeta anticlines in
constraints in the time– temperature (t-T ) space derived
from stratigraphical relationships in the well. Thermal the western foothills) adjacent to the synclinal mar-
solutions support the onset of cooling in Palaeocene time. ginal depocentres in both the western and eastern
foothills (Medina, Nunchı́a in the east, and Nuevo
Mundo and Guaduas in the west) (Fig. 2) were
Horton et al. 2010a; Mora et al. 2010b, 2011). New exhumed by thrust-induced denudation, while
exhumation data in Silva et al. (this volume, in growth strata in adjacent synclines underwent defor-
press) confirms that this period recorded the mation in which sedimentation rates mostly
maximum frontal advance of orogenic deformation exceeded crestal uplift rates of the main structures
(e.g. Parra et al. 2009b). The dramatic eastwards (Fig. 12). Even more important is the finding that
advance of the pinch-out of the Oligocene C7 exhumation of these basement highs after the
member of the Carbonera Formation (Fig. 11a) period of tectonic quiescence was typically coeval
documented with our well and subsurface data (see the References cited above), suggesting a syn-
reinforces this interpretation. Effectively, the zero chronous activation of the main syncline-inversion
depositional boundary in the foreland was pushed anticline pairs.
eastwards by the advancing deformation front.
Microtectonic data (Mora et al. 2013) also show
that second-order detachment and basement folds Acceleration in tectonic rates and
formed in the Oligocene in different places of the subsequent peak deformation
Eastern Cordillera with fractures and associated
slaty cleavage of that age. These folds were gener- We show that the Early Miocene was a period with
ally not associated with inversion basement faults, comparatively higher accumulation rates in both
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 9. Cenozoic sediment accumulation rates based on both compacted (CAR) and decompacted (DAR) thicknesses
in the NE Magdalena Valley (Nuevo Mundo), and northern (Sirirı́) and southern (Medina) eastern foothills. Data
sources are Parra et al. (2010).

Fig. 10. Double-dated ZHe ages of samples from the Floresta area in the Eastern Cordillera plotted by their stratigraphic
age and lag time (e.g. Ruiz et al. 2004, dashed diagonal lines) (after Saylor et al. 2012). Zircons are identified as being of
volcanic origin if their ZHe and Zircon U– Pb ages overlap within their 2s uncertainty. Volcanic zircons (red) are
excluded from lag-time analysis. The three stages are interpreted as episodes of rapid exhumation (stages 1 and 2) and
the introduction of new supra-PRZ sedimentary sources (Stage 3). Lag-time values (L) are in Ma.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 11. Oligocene–Miocene palaeogeography of the study area based on surface and subsurface cross-cutting
stratigraphical relationships, and thermochronometric and provenance data. See Caballero et al. (2013) and Silva et al.
(this volume, in press), and references therein, for supporting information.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 12. Subsurface cross-cutting relationships indicating Oligocene deformation. Oligocene growth strata occur in the
(a) central and (b) northern eastern foothills, and (c) in the Provincia oil field in the Middle Magdalena Basin. The
location of seismic sections is provided in Figure 2.

proximal basins to the east and west of the Eastern In addition, direct qualitative evidence suggests
Cordillera (Fig. 9). The coarser-grained record of that shortening rates should have beeen faster
the C1 member in the Medina Basin (Parra et al. since the Late Miocene in both foothill areas. This
2010), as well as in the Magdalena (Colorado For- is supported by the following observations.
mation: Gómez et al. 2005; Moreno et al. 2011;
Caballero et al. this volume, in press), may sug- † Cross-cutting relationships in cross-sections
gest increased rates of deformation. Indeed, internal from both foothill belts, especially evident
unconformities within the Colorado Formation in the northernmost cross-sections (Jiménez
appear to support that idea (see Fig. 12a). et al. this volume, in press) where the longest
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 12. Continued.

traces of the most-frontal faults cut through Late upper portion of the Guayabo Formation should
Miocene– Pliocene continental sediments of the be Pleistocene. If true, then cross-sections L and
Guayabo Formation. M from Jimenez et al. (this volume, in press)
† Thermochronology reported by Mora et al. would require virtually all of the approximately
(2008, 2010b, c), as well as Moreno et al. (this 10 km of shortening along the Yopal Thrust to be
volume, in press), from the most-frontal faults of Late Miocene –Pleistocene age, with more than
in both foothills suggest the fastest deformation half of that as Pleistocene. Therefore, shortening
rates occurred along those deformation fronts. rates of at least 3 mm year21 should be obtained
for those frontal thrusts. These are in line with the
To consider the detailed quantification of these approximately 5 mm year21 rate obtained by Mora
rates, it is interesting to review the recent chro- et al. (2008) in the southern segment of the eastern
nology of the Neogene Guayabo Formation by foothills.
Delgado et al. (2012). Using AFT data and intra- New U– Pb ages from the Real Formation along
Guayabo unconformities, they suggest that the the western foothills (Caballero et al. 2013) give
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 12. Continued.

sedimentation rates of about 0.35 mm year21 in the pattern is detected with little Neogene exhumation
Late Miocene (Fig. 13) compared with rates in the central axial zone near Bogotá (samples
of 0.25 –0.40 mm year21 in the eastern foothills RG06, RON1002 and US1004: Fig. 15) contrast-
(Fig. 9). Sedimentation rates clearly increased at ing with the external deeply incised valleys in the
the same time as shortening rates in both areas. How- foothills where exhumation accelerated drasti-
ever, of note in our cross-sections (see also Tesón cally (1023– 24, MM005, CS3). In the Cocuy area,
et al. this volume, in press) is the presence of a models from the northern part of the eastern foot-
thicker Neogene (especially Late Miocene –Pleisto- hills also require fast exhumation rates since the
cene) sedimentary sequence in the eastern foothill Late Miocene (sample AFT 250710-12: Fig. 15).
belt. This feature is especially visible in the north- A similar pattern of rapid Late Miocene –Recent
ernmost regional cross-section with the thickest suc- exhumation is observed to the north in the Sirirı́
cession (Fig. 14). area, where AFT ages younger than about 5 Ma
An additional aspect becomes evident when occur and the Precambrian basement has been
one compares the Neogene exhumation patterns deeply incised (Fig. 16). This is in line with the
obtained with multiple thermochronometers mod- observations by Mora et al. (2008) of faster exhuma-
elled using HeFTy software (Ketcham 2005) in tion rates in the foothills during the Neogene. If our
different geomorphic areas of the eastern flank of initial shortening assessments are considered, then
the Eastern Cordillera. These data are from the a coeval acceleration of shortening would be also
incised valleys of the eastern foothills, which con- required in the Neogene. Therefore, we try in the
trast with the very low incision of almost flat-lying following section to give more precision to our
areas in the axial zone near Bogotá. A distinct shortening assessments throughout the Cenozoic
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 13. Field relationships (a) and detrital zircon U–Pb ages, shown in (b) as age –probability functions (red line)
and age histograms (blue), in three Miocene samples of the Middle–Upper Miocene Real Formation in the Opon
Anticline, Middle Magdalena Valley Basin. Middle– Late Miocene ages of the youngest populations of detrital zircon
U– Pb ages overlap with biostratigraphical and radiometric ages of the unit (see Gómez et al. 2005 and references
therein) and exhibit an upsection decrease in age. This pattern reflects the volcanoclastic origin of zircons and provides
maximum depositional ages. Thickness relationships (inset cross-section in a) and ages yield compacted sediment
accumulation rates of approximately 0.35 mm year21. See the location in Figure 2.

by kinematically restoring the deformed-state cross- Silva et al. this volume, in press), in addition to
sections of Tesón et al. (this volume, in press) (Figs published work (Parra et al. 2009b; Mora et al.
14, 17–19). 2010b; Ramı́rez-Arias et al. 2012). Later, we
compare these results with an assessment of the
Extrapolation of shortening rates from the rocks exposed at the surface from various prove-
nance studies in the same area (Caballero 2010;
regional balanced cross-sections
Quintero 2010; Parra et al. 2010; Moreno et al.
We used a procedure where all of the data on the 2011; Ramı́rez-Arias et al. 2012), which allow us
chronology of deformation is incorporated into the to calibrate the kinematics of fault movement
kinematic restoration. Ideally, this includes the use given a prescribed geometry.
of growth strata. However, such records are particu- From this approach, in all sections, we obtained
larly scarce in the Eastern Cordillera. Therefore, as Palaeocene shortening rates of less than 2 mm
described previously, we mostly relied on deriving year21, which then decreased to almost zero during
the position of buried rocks through time from the mid-Eocene and then started increasing again
the modelled thermal histories and a geothermal by the Oligocene (Fig. 20). However, since the
gradient using the raw data and the HeFTy models Oligocene, shortening rates have been 2 mm
in this volume (Caballero et al. 2013, this volume, year21 before 10 Ma, with a dramatic increase after-
in press; Moreno et al. this volume, in press; wards, which peaks at values of between 5 and
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 14. Sequentially retrodeformed structural cross-section from the Middle Magdalena Valley, across the Santander
Massif to the NE foothills, approximately at 78N. The location of the section is provided in Figure 2.

8 mm year21 in the regional sections, and between the Palaeocene was mostly focused in the Central
3 and 5 mm year21 for the eastern and western foot- Cordillera, the Magdalena Valley and in the wes-
hills of the Eastern Cordillera. Cross-sections with ternmost portions of the Eastern Cordillera, as evi-
sufficient resolution for different steps during Late denced by the exhumation records of those areas
Miocene–Pleistocene time provide evidence for a (Fig. 5a, b). However, local and presumably low-
gradual acceleration in shortening rates from the amplitude structural highs have been also documen-
moderate Late Miocene rates to the very fast Plio- ted in eastern areas as far as the Llanos Basin
Pleistocene (Fig. 20). (Fig. 6). Documented rates of shortening would be
less than 1 mm year21 (Fig. 20). We suggest that
this period involved localized and strong defor-
Discussion mation in the Central Cordillera but widely spaced
and mild inversion in the Eastern Cordillera and
In order to propose a model for the observed spatio- Llanos Basin. In contrast, after deformation
temporal deformation patterns, in the following resumed in the Latest Eocene (Fig. 5c, d), exhuma-
discussion we suggest a plausible correlation of tion during the Oligocene was documented all along
different deformation processes with the plate the former Early Cretaceous graben, an area that
tectonic configuration, mechanical properties of coincides with the present-day positive relief of
deformed rocks and the role of surface processes. the Eastern Cordillera (Fig. 11a). Shortening rates
were also faster and, locally, reached 2 mm year21,
The role of plate tectonics coeval with the fastest eastwards advance of the oro-
genic front. At this time, the easternmost Eastern
It is important to recognize that the spatial pattern of Cordillera chain should have reached areal dimen-
deformation clearly shows that shortening during sions similar to today but, presumably, at lower
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 14. Continued.

elevation (Fig. 11a). Peak shortening rates and out- Oligocene –Neogene shortening and the increasing
of-sequence reactivation of the main inversion rates of westwards South American drift (Silver
faults were in place, starting in latest Miocene time, et al. 1998). In such a case, the shortening threshold
during a phase of topographical growth and eleva- values of about 2 cm year21 for westwards South
tion rise (Van der Hammen et al. 1973; Wijninga American drift in the Eastern Cordillera would be
1996; Hooghiemstra et al. 2006) in an inversion slightly higher than those in the Central Andes
orogen that already attained its broadest areal since deformation also started later in the Eastern
extent in the Oligocene (Fig. 11). Cordillera. However, it is worth noting that post-
In the following, we try to correlate the observed Early Eocene shortening could have commenced
deformation behaviour with Cenozoic plate kin- slightly earlier in the Central and Western Cordil-
ematics. A phase of shortening, starting at about leras to the west. In this context, the most-important
40 Ma, with subsequently increasing rates is dif- phase of elevation increase in the Eastern Cordil-
ficult to correlate with Nazca – South America con- lera would correlate with the faster Late Miocene
vergence rates. For instance, a peak conver- shortening rates in the upper plate and the fastest
gence rate is defined at approximately 49 Ma and westward South American drift.
a minimum by about 40 Ma (Pardo-Casas & Another interesting correlation is the width of
Molnar 1987). This clearly anti-correlates with the the deformation zone across the Early Cretaceous
shortening rates we report here for the Northern graben domain of the Eastern Cordillera and the
Andes. plate kinematic history. First, we find that by
In the Central Andes, Oncken et al. (2006) 49 Ma, published plate kinematic data (Pardo-Casas
estimated that when rates of westward South Amer- & Molnar 1987) (Fig. 21) suggest that the direction
ican drift reach threshold values of about 1.7 cm of convergence becomes more perpendicular with
year21, shortening starts in the upper plate. As in respect to the South American margin and remains
the Central Andes, with our interpretation of relatively uniform during the following periods.
Eastern Cordillera shortening resuming again by We suggest that more localized deformation in the
the Late Eocene (c. 40 Ma), we find a rough coinci- Palaeocene is caused by a more oblique to transpres-
dence between the onset, then steadily increasing sional regime that did not facilitate deformation to
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013
A. MORA ET AL.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

advance eastwards towards the weak graben domain master inversion faults with opposite vergence in
but, instead, promoted deformation along deeply both the western and eastern foothills continued to
rooted exhuming structures. However, the wide- be active and shed detritus into adjacent Cenozoic
spread low-amplitude deformation in the Llanos depocentres. These depocentres are bounded by
Basin could be due to a preferential reactivation of either shortcuts or Neocomian faults affected by
local inversion structures under low strain rates. In flexural normal faulting (Caballero et al. 2013;
this context, low strain rates favour a more wide- Moreno et al. this volume, in press). We propose
spread, less focused mode of deformation. In con- that this association of coeval activation of base-
trast, the very fast shortening rates in the late ment highs and adjacent slower-slip shortcuts
Miocene, when the plates converged in a near- should be a fundamental behaviour of inverted
perpendicular direction, would favour focused symmetric grabens, which starkly contrasts with
plastic or brittle failure of the crust, which may acti- other thrust belts. Thus, in this inverted orogen,
vate deeply rooted master normal faults and prompt there are faults of similar mechanical strength
increased elevation (Figs 13– 16). for reactivation but opposite vergence, and we
show that these became active at the same time on
Mechanical properties both sides of the orogen and along strike in the
same front (Mora et al. 2006, 2010b; Parra et al.
As previously mentioned, the metasedimentary 2009b; Bande et al. 2012; Ramı́rez-Arias et al.
basement in the Eastern Cordillera compared with 2012; Caballero et al. 2013; Moreno et al. this
the cratonic basement of the Llanos and Putumayo volume, in press).
areas or Maracaibo Basin, along with the nearly This general behaviour, as well as the regional
absent shortening in these two cratonic areas, spatial pattern of propagation of basement-involved
suggest the presence of a long-lasting weak mobile deformation, suggests that deformation partition-
belt between two strong cratonic blocks. This model ing with time is controlled by the former rift sub-
supports the indenter model (Taboada et al. 2000), graben and master faults (Fig. 4). However, before
where the weak block is squeezed by the two resist- reactivation and brittle faulting can occur, thus
ant ones and moves along the Algeciras and Bucar- producing different styles (described in detail by
amanga faults. However, assessments by Tesón Tesón et al. this volume, in press), strain hardening
et al. (this volume, in press) suggest that the is required. The main evidence for this conclusion is
weakest domain in the Eastern Cordillera is prob- the presence of Oligocene second-order folds (i.e.
ably the one where contractional deformation is folds that are independent of inversion tectonics)
intense during the Late Palaeozoic (e.g. Permo- and associated coeval penetrative strain (Mora
Triassic metamorphism west of the Boyacá and et al. 2013). The coeval activation of inherited
Bucaramanga faults: see Fig. 4c) and Mesozoic sub- fault systems in inverted symmetric grabens should
sidence first commences (e.g. documented Jurassic be a typical feature for this style of orogenic belts
Girón Formation: Kammer & Sánchez 2006). This since it has been observed even in domains where
point is reinforced by the fact that deformation the reported stress directions exhibit different direc-
associated with significant cooling starts at, and is tions with respect to the main faults in other domains
restricted to, this precise domain during the Palaeo- of the belt.
cene, coinciding with the Early Cretaceous graben Certainly, the northwards escape of the weak
basin. Therefore, under low shortening rates (dur- block, as suggested by Tesón et al. (this volume,
ing the Palaeocene) only the weakest portions of in press), causes oroclinal bending of the Eastern
the graben underwent significant deformation. In Cordillera mobile belt between the two cratonic
contrast, faster Oligocene shortening rates (and a areas. Although such lateral changes in the pre-
more perpendicular convergence) prompted a existing configuration promote salient growth in the
virtually coeval activation of all of the inherited orogen, this does not necessarily indicate an earlier
master graben faults (see Caballero et al. 2013; or later activation of the master basin-boundary
Moreno et al. this volume, in press; Tesón et al. graben faults during contraction. Rather, these
this volume, in press), which slipped at rates results most probably suggest that basin-boundary
higher than the adjacent footwall shortcuts (see master normal faults of opposite vergence (Magda-
Figs 14 & 17 –19). During the Oligocene, analogue lena and eastern foothills) somehow interacted

Fig. 15. Contrasting thermal histories based on AFT and (U– Th)/He data from rocks cropping out along the axial
sector and along the eastern foothills of the Eastern Cordillera. Late Eocene–Oligocene (40– 30 Ma) onset of cooling
and moderate to low cooling rates since late Miocene time along the axial Eastern Cordillera contrast with Late
Oligocene– early Miocene (30–20 Ma) onset of cooling and very rapid Late Miocene– Recent cooling rates in the
eastern foothills. See Figure 8 for conventions for thermal models plots.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 16. (a) Geological map and AFT ages from the Sirirı́ area in the NE foothills of the Eastern Cordillera. (b) Thermal
models from selected samples. See Figure 8 for conventions.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

(a)

(b)

Fig. 17. Sequentially retrodeformed structural cross-section from the Opon Anticline in the western foothills,
across the Floresta Massif, to the Zamaricote Syncline in the central eastern foothills, approximately at 68 N. See
the location in Figure 2.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

(a)

(b)

Fig. 18. Sequentially retrodeformed structural cross-section from the La India Syncline in the western foothills, across
the northern Bogotá Plateau, to the Nunchı́a Syncline in the central eastern foothills, approximately at 5810′ N.
See the location in Figure 2.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Fig. 19. Sequentially retrodeformed structural cross-section approximately at 48N, from the southern Guaduas
Syncline in the western foothills, across the Bogotá Plateau, to the Boa Anticline in the SE foothills. See the location in
Figure 2.

in the extensional phase, transferring stresses from particular shortening rates (roughly 1–2 mm
one side of the graben to the other in the absence year21), the system is again activated as a whole
of hard linkages. Later in the Oligocene, under with interacting faults.

Fig. 20. (a) Shortening rates calculated from sequentially restored cross-sections presented in Figures 14 and 17– 19.
East– west-spreading velocity of the South America–Africa (bold black line, after Silver et al. 1998) is provided for
comparison. (b)Shortening rates in local cross-sections in the western (Lisama) and eastern (Sirirı́ and
Volcanera) foothills.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

Fig. 21. Reconstruction of two points on the Nazca Plate with respect to the South American Plate at the times of
various magnetic anomalies, after Pardo-Casas & Molnar (1987). A major change in convergence obliquity
at Early Eocene time (49 Ma) is indicated with a blue ellipse.

Surface processes relief. Regarding shortening rates, faster rates


coupled with rapid exhumation is something not
As mentioned in Nemcŏk et al. (this volume, in necessarily observed in the eastern side of the
press) and previously discussed in the works by Eastern Cordillera. Indeed, the fastest rates of short-
Mora et al. (2008) and Ramı́rez-Arias et al. ening for the western and eastern sides started after
(2012), we reinforce the idea that the Eastern Cor- the Latest Miocene and continued to present. In this
dillera is a key natural laboratory for studying case we do not find conclusive new evidence of
inverted orogens conditioned by climatic forcing. the eastern side of the Eastern Cordillera moving
In the Eastern Cordillera, focused denudation due faster than the western side due to more rapid
to an orographic rain shadow on its eastern side is denudation.
a typical feature. Climatic forcing prompted by Faster denudation rates, of the order of
moisture-bearing winds coming from the Amazon 1 mm year21 (Mora et al. 2008), certainly produce
apparently focused precipitation, differentially other effects. Most obvious is the massive flux of
enhancing denudation, as we show here, and also Neogene –Recent conglomerates during periods of
unloading the active faults (Mora et al. 2008). peak accumulation (Fig. 9). With the available
Towards the zone of lateral escape (Cocuy area) in data, this appears to be limited to the eastern foot-
the weak block of the Eastern Cordillera, the Cordil- hills, because the western Magdalena foothills have
lera not only advanced through a salient, but it also thinner Neogene –Recent conglomeratic sequences
developed extremely high structural and topogra- and slower accumulation rates. We hypothesize
phical relief associated with very young AFT and that rapid fault movement and synchronous rapid
AHe ages (Fig. 16). This is regarded as evidence erosion generate a massive flux of sediment that
of peak denudation rates, which facilitate kinematic is exclusive to the eastern foothills. As suggested
histories in which structural blocks prefer to grow by Jiménez et al. (this volume, in press), this thick
vertically and not advance. Additional evidence sequence of Neogene siliciclastic rocks acted as a
for this point of view is that the weakest central rigid backstop during deposition of the uppermost
block of the mobile belt, west of the Boyacá levels and also heated deeper shaly horizons,
and Bucaramanga faults, which exhumed deep prompting them to reduce basal friction. In certain
metamorphic basement during late Palaeozoic oro- areas of the eastern foothills, with all of these
genesis and is adjacent to upper Palaeozoic sedi- elements, antiformal stacks and passive roof
mentary rocks, was not the locus of deep or duplexes are present. It is significant how this type
focused exhumation during Tertiary contraction of structure is underdeveloped or absent in the Mag-
and neither focused deformation and structural dalena foothills where the Neogene clastics are
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

thinner and the basal shaly detachments are more Bayona, G., Valencia, A., Mora, A., Rueda, M., Ortiz,
discontinuous (see Moreno et al. this volume, in J. & Montenegro, O. 2008. Estratigrafı́a y pro-
press). The combination of these factors has been cedencia de las rocas del Mioceno en la parte distal
reproduced with similar results in analogue models de la Cuenca Llanos. Geologı́a Colombiana, 33,
23–46.
(Couzens-Schultz et al. 2003). Bayona, G., Cardona, A. et al. 2013. Onset of fault
reactivation in the Eastern Cordillera of Colombia
and proximal Llanos Basin; response to Caribbean–
Conclusions South American convergence in early Palaeogene
time. In: Nemčok, M., Mora, A. & Cosgrove, J. W.
The Eastern Cordillera of Colombia is a prime (eds) Thick-Skin-Dominated Orogens: From Initial
example of inversion tectonics where plenty of Inversion to Full Accretion. Geological Society,
new data sets allow us to evaluate the role of differ- London, Special Publications, 377, first published
ent factors in mountain building in an inverted rift online March 8, 2013, http://dx.doi.org/10.1144/
setting. SP377.5
Butler, R. W. H. 1987. Thust sequences. Journal of the
† Plate tectonics, mainly the direction of conver- Geological Society, 144, 619–634.
gence and the velocity of westwards South Caballero, V. 2010. Evolución tectono-sedimentaria
American drift most probably controlled the del Sinclinal de Nuevo Mundo, Cuenca Valle Medio
width of the deformation zones and the defor- del Magdalena Colombia, durante el Oligoceno–
mation rates. Mioceno. MSc thesis, Universidad Industrial de
† The presence of a weak, long-lasting mobile belt Santander.
that controls multiphased long-lived tectonic Caballero, V., Mora, A. et al. In press. Tectonic con-
events defines the Eastern Cordillera. Of impor- trols on sedimentation in an intermontane hinterland
basin adjacent to inversion structures: the Nuevo
tance is the realization that the contractional
Mundo Syncline, Middle Magdalena Valley, Colom-
reactivation of most of its main inversion base- bia. In: Nemčok, M., Mora, A. & Cosgrove, J. W.
ment highs is virtually coeval regardless of its (eds) Thick-Skin-Dominated Orogens: From Initial
vergence. Subdivision of the region into weaker Inversion to Full Accretion. Geological Society,
blocks is also possible, as represented by areas London, Special Publications, 377, http://dx.doi.org/
where the earliest (Mesozoic) extension and 10.1144/SP377.12
the earliest (Palaeozoic) contraction occurred Caballero, V., Parra, M., Mora, A., Lopez-Arias, C.,
within the broader domain of the Mesozoic Rojas, L. E., Quintero, I. & Horton, B. K. 2013.
graben basin (west of the Bucaramanga and Factors controlling selective abandonment and reacti-
vation in thick skin orogens: A case study in the Mag-
Boyacá faults). dalena Valley, Colombia. In: Nemčok, M., Mora,
However, surface processes facilitated focused pro- A. & Cosgrove, J. W. (eds) Thick-Skin-Dominated
tracted denudation and allowed deformation to be Orogens: From Initial Inversion to Full Accretion.
focused in the areas on the lee side of the orogen, Geological Society, London, Special Publications,
but generally not in the weakest zones 377, first published online March 8, 2013, http://dx.
doi.org/10.1144/SP377.4
This study was funded by Ecopetrol–ICP as part of the Casero, P., Salel, J. F. & Rossato, A. 1997. Multidisci-
project ‘Cronologı́a de la Deformación en las Cuencas plinary correlative evidences for polyphase geologi-
Subandinas’. The authors would like to thank the editors cal evolution of the foot-hills of the Cordillera
of this Special Publication, J. Cosgrove and M. Nemcŏk, Oriental (Colombia). In: VI Simposio Bolivariano de
for their constructive comments; and also the reviewers Exploración Petrolera en las Cuencas Subandinas.
J. Buchanan, A. Henk and R. Sorkhabi for their positive Bogotá, Colombia. ACGGP (Asociación Colombiana
contributions that improved the manuscript. H. Bueno was de Geólogos y Geofı́sicos del Petroleo), Bogotá,
of great help with the final edition of the cross-sections. 100– 118.
Cooper, M. A., Addison, F. T. et al. 1995. Basin devel-
opment and tectonic history of the Llanos Basin,
References Eastern Cordillera, and Middle Magdalena Valley,
Colombia. AAPG Bulletin, 79, 1421– 1443.
Allmendinger, R. W., Jordan, T. E., Kay, S. M. & Cortés, M. & Angelier, J. 2005. Current states of stress
Isacks, B. L. 1997. The evolution of the Altiplano-Puna in the northern Andes as indicated by focal mechan-
Plateau of the Central Andes. Annual Review of Earth isms of earthquakes. Tectonophysics, 403, 29–58.
and Planetary Sciences, 129, 139–144. Cortés, M., Colletta, B. & Angelier, J. 2006. Struc-
Bande, A., Horton, B. K., Ramı́rez-Arias, J. C., Mora, ture and tectonics of the central segment of the
A., Parra, M. & Stockli, D. 2012. Clastic deposition Eastern Cordillera of Colombia. Journal of South
and detrital provenance of evolving sediment source American Earth Sciences, 21, 437 –465.
regions in the frontal Eastern Cordillera, Colombia: Couzens-Schultz, B. A., Vendeville, B. C. &
Implications for the sequence of Andean thrust defor- Wiltschko, D. V. 2003. Duplex style and triangle
mation. Geological Society of America Bulletin, 124, zone formation: insights from physical modeling.
59– 76, http://dx.doi.org/10.1130/B30412.1 Journal of Structural Geology, 25, 1623– 1644.
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

DeCelles, P. G. & Horton, B. K. 2003. Early to middle Orogen of Amazonia and its implications for Rodinia
Tertiary foreland basin development and the history of reconstructions: new U– Pb geochronological insights
the Andean crustal shortening in Bolivia. Geological sinto the Proterozoic tectonic evolution of northwes-
Society of America Bulletin, 115, 58–77. tern South America. Precambrian Research, 191,
DeCelles, P. G. & Mitra, G. 1995. History of the Sevier 58–77, http://dx.doi.org/10.1016/j.precamres.2011.
orogenic wedge in terms of critical taper models, 09.005.
northeast Utah and southwest Wyoming. Geological INGEOMINAS & Mora, A. 2005. Levantamiento de
Society of America Bulletin, 107, 454–462. información estratigráfica y estructural de los cintur-
Delgado, A., Mora, A. & Reyes-Harker, A. 2012. ones esmeraldı́feros de la Cordillera Oriental. Instituto
Deformation partitioning in the Llanos foreland basin Colombiano de Geologı́a y Mineria (Ingeominas),
during the Cenozoic and its correlation with defor- Bogotá.
mation in the hinterland. Journal of South American Jiménez, L., Mora, A. et al. In press. Segmentation and
Earth Sciences, 39, 228– 244. growth of foothill thrust belts adjacent to inverted
Echavarria, L., Hernández, R., Allmendinger, R. & grabens: the case of the Colombian eastern foothills.
Reynolds, J. 2003. Subandean thrust and fold belt of In: Nemčok, M., Mora, A. & Cosgrove, J. W.
northwestern Argentina: Geometry and timing of the (eds) Thick-Skin-Dominated Orogens: From Initial
Andean evolution. AAPG Bulletin, 87, 965– 985. Inversion to Full Accretion. Geological Society,
Escalona, A. & Mann, P. 2006. Tectonic controls on the London, Special Publications, 377, http://dx.doi.org/
rigth-lateral Burro Negro tear fault on Paleogene struc- 10.1144/SP377.11
ture and stratigraphy, northeastern Maracaibo Basin. Johnsson, M. J., Stallard, R. F. & Lundberg, N. 1991.
AAPG Bulletin, 90, 479–504. Controls on the composition of fluvial sands from a tro-
Farris, D. W., Jaramillo, C. et al. 2011. Fracturing the pical weathering environment: Sands of the Orinoco
Panamanian Isthmus during initial collision with South River drainage basin, Venezuela, Colombia. Geologi-
America. Geology, 39, 1007–1010, http://dx.doi.org/ cal Society of America Bulletin, 103, 1622– 1647.
10.1130/G32237.1. Kammer, A. 1997. Los pliegues del sinclinal de Tunja:
Galbraith, R. F. 1981. On statistical models for fission análisis estructural y modelamiento geométrico. Geo-
track counts. Journal of the International Association logı́a Colombiana, 22, 3 –25.
for Mathematical Geology, 13, 471– 478. Kammer, A. & Mora, A. 1999. Structural styles of the
Gómez, E., Jordan, T. E., Allmendinger, R. W., folded Bogotá Segment, Eastern Cordillera, Colombia.
Hegarty, K., Kelley, S. & Heizler, M. 2003. Zentralblatt für Geologie und Paleontologie, Teil 1,
Controls on architecture of the Late Cretaceous to Cen- 823–837.
ozoic southern Middle Magdalena Valley Basin, Colom- Kammer, A. & Sánchez, J. 2006. Early Jurassic rift struc-
bia. Geological Society of America Bulletin, 115, tures associated with the Soapaga and Boyacá Faults of
131–147. the Eastern Cordillera, Colombia: sedimentological
Gómez, E., Jordan, T. E., Allmendinger, R. W., inferences and regional implications. Journal of
Hegarty, K. & Kelley, S. 2005. Syntectonic Ceno- South American Earth Sciences, 21, 412–422.
zoic sedimentation in the northern middle Magdalena Ketcham, R. A. 2005. Forward and inverse modeling of
Valley Basin of Colombia and implications for exhu- low-temperature thermochronometry data. Reviews in
mation of the Northern Andes. Geological Society of Mineralogy and Geochemistry, 58, 275– 314.
America Bulletin, 117, 547– 569. Mann, P. & Vargas-Jimenez, C. A. 2011. Role of
Green, P. F. 1981. “Track-in-track” length measurements Panama arc indentor for late Cenozoic deformation
in annealed apatites. Nuclear Tracks, 5, 121–128. in Colombia and implications for regional distribution
Gutscher, M. A., Malavieille, J., Lallemand, S. & of hydrocarbons, AAPG Annual Convention and Exhi-
Collot, J. Y. 1999. Tectonic segmentation of the bition, Houston, Texas, USA.
North Andean margin: impact of the Carnegie Ridge Mora, A. & Kammer, A. 1999. Comparación de los estilos
Collision. Earth and Planetary Science Letters, 168, estructurales en la sección entre Bogotá y los Faral-
255– 270. lones de Medina, Cordillera Oriental de Colombia.
Hooghiemstra, H., Wijninga, V. M. & Cleef, A. M. Geologı́a Colombiana, 24, 55–82.
2006. The Paleobotanical record of Colombia: impli- Mora, A., Parra, M., Strecker, M. R., Kammer, A.,
cations for biogeography and biodiversity. Annals of Dimaté, C. & Rodriguez, F. 2006. Cenozoic con-
the Missouri Botanical Garden, 93, 297– 325. tractional reactivation of Mesozoic extensional
Horton, B. K., Parra, M., Saylor, J. E., Nie, J., Mora, structures in the Eastern Cordillera of Colombia. Tec-
A., Stockli, D. F. & Strecker, M. 2010a. Resolving tonics, 25, TC2010, http://dx.doi.org/10.1029/2005
uplift of the northern Andes using detrital zircon age TC001854.
signatures. GSA Today, 20, 4 –9. Mora, A., Parra, M., Strecker, M. R., Sobel, E. R.,
Horton, B. K., Saylor, J. E., Nie, J., Mora, A., Parra, Hooghiemstra, H., Torres, V. & Vallejo-
M., Reyes-Harker, A. & Stockli, D. F. 2010b. Jaramillo, J. 2008. Climatic forcing of asymmetric
Linking sedimentation in the northern Andes to base- orogenic evolution in the Eastern Cordillera of Colom-
mento configuration, Mesozoic extension, and Ceno- bia. Geological Society of America Bulletin, 120,
zoic shortening: Evidence from detrital zircon U–Pb 930–949.
ages, Eastern Cordillera, Colombia. Geological Soci- Mora, A., Baby, P., Roddaz, M., Parra, M., Brusset, S.,
ety of America Bulletin, 122, 1423– 1442. Hermoza, W. & Espurt, N. 2010a. Tectonic history
Ibañez-Mejia, M., Ruiz, J., Valencia, V. A., Cardona, of the Andes and sub-Andean zones: implications for
A., Gehrels, G. & Mora, A. 2011. The Putumayo the development of the Amazon drainage basin. In:
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

INVERSION TECTONICS: COLOMBIAN ANDES

Hoorn, C. & Wesselingh, F. P. (eds) Amazonia, controls on Cenozoic foreland basin development in
Landscape and Species Evolution, 1st edn. Blackwell, the north-eastern Andes, Colombia. Basin Research,
Oxford. 22, 874 –903.
Mora, A., Horton, B. K. et al. 2010b. Migration of Cen- Parra, M., Mora, A., Lopez, C., Rojas, L. E. & Horton,
ozoic deformation in the eastern Cordillera Colombia B. K. 2012. Detecting earliest shortening and defor-
interpreted from fission track results and structural mation advance in thrust-belt hinterlands: example
relationships: Implications for petroleum systems. from the Colombian Andes. Geology, 40, 175– 178,
AAPG Bulletin, 94, 1543– 1580. http://dx.doi.org/10.1130/G32519.1.
Mora, A., Parra, M. et al. 2010c. The eastern foothills Paton, D. A. & Underhill, J. R. 2004. Role of crustal
of the Eastern Cordillera of Colombia: an example of anisotropy in modifying the structural and sedimento-
multiple factors crontolling structural styles and logical evolution of extensional basins: the Gamtoos
active tectonics. Geological Society of America Bulle- Basin, South Africa. Basin Research, 16, 339 –359.
tin, 122, 1846, http://dx.doi.org/10.1130/B30033.1. Quintero, I. 2010. Cinemática de las estructuras de
Mora, A., Blanco, V. et al. 2013. On the lag time deformación de la Cuenca de Medina (Colombia)
between internal strain and basement involved thrust durante la depositacion del Grupo Guayabo. MSc
induces exhumation: The case of the Colombian thesis, Universidad Industrial de Santander.
Eastern Cordillera. Journal of Structural Geology, in Ramı́rez-Arias, J. C., Mora, A., Rubiano, J., Duddy, I.,
press. Parra, M., Moreno, N., Stockli, D. & Casallas, W.
Moreno, C. J., Horton, B., Caballero, V., Mora, A., 2012. The asymetric evolution of the Colombian
Parra, M. & Sierra, J. 2011. Depositional and prove- Eastern Cordillera. Tectonic inheritance or climatic
nane record of the Paleogene transition from foreland forcing? New evidence from thermochronology and
to hinterland basin evolution during Andean orogen- sedimentology. Journal of South American Earth
esis, northern Middle Magdalena Valley basin, Colom- Sciences, 39, 112 –137, http://dx.doi.org/10.1016/
bia. Journal of South American Earth Sciences, 32, j.jsames.2012.04.008.
246–263, http://dx.doi.org/10.1016/j.jsames.2011. Ring, U. 1994. The influence of preexisting structure on
03.018. the evolution of the Cenozoic Malawi rift (East
Moreno, N., Silva, A. et al. In press. Interaction African rift system). Tectonics, 13, 313– 326.
between thin- and thick-skinned tectonics in the foot- Ruiz, G. M. H., Seward, D. & Winkler, W. 2004. Detri-
hill areas of an inverted graben. The Middle Magda- tal thermochronology – a new perspective on hinter-
lena Foothill belt. In: Nemčok, M., Mora, A. & land tectonics, an example from the Andean Amazon
Cosgrove, J. W. (eds) Thick-Skin-Dominated Oro- Basin, Ecuador. Basin Research, 16, 413–430.
gens: From Initial Inversion to Full Accretion. Geo- Sarmiento-Rojas, L. F., Van Wess, J. D. & Cloetingh,
logical Society, London, Special Publications, 377, S. 2006. Mesozoic transtensional basin history of the
http://dx.doi.org/10.1144/SP377.18 Eastern Cordillera, Colombian Andes: Inferences
Nemcŏk, M., Hermeston, S. et al. In press. Thick-skin from tectonic models. Journal of South American
orogens – foreland interactions and their controlling Earth Sciences, 21, 383–411.
factors, Northern Andes of Colombia. In: Nemčok, Saylor, J. E., Stockli, D. F., Horton, B. H., Nie, J. &
M., Mora, A. & Cosgrove, J. W. (eds) Thick-Skin- Mora, A. 2012. Discriminating rapid exhumation
Dominated Orogens: From Initial Inversion to Full from syndepositional volcanism using detrital zircon
Accretion. Geological Society, London, Special Publi- double dating: Implications for the tectonic history of
cations, 377, http://dx.doi.org/10.1144/SP377.16 the Eastern Cordillera, Colombia. Geological Society
Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P. of America Bulletin. 124, 762– 779, http://dx.doi.
& Schemmann, K. 2006. Deformation of the Cen- org/10.1130/B30534.1
tral Andean upper plate system – facts, fiction, and Silva, A., Mora, A. et al. In press. Basin compartmenta-
constraints for plateau models. In: Oncken, O., lization and drainage evolution during rift positive
Chong, G. et al. (eds) The Andes: Active Subduction inversion: evidence from multiple techniques in the
Orogeny. Frontiers in Earth Sciences. Springer, Berlin, Eastern Cordillera of Colombia. In: Nemčok, M.,
3–27. Mora, A. & Cosgrove, J. W. (eds) Thick-Skin-
Pardo-Casas, F. & Molnar, P. 1987. Relative motion of Dominated Orogens: From Initial Inversion to Full
the Nazca (Farallon) and South American plates since Accretion. Geological Society, London, Special Publi-
Late Cretaceous time. Tectonics, 6, 233–248. cations, 377, http://dx.doi.org/10.1144/SP377.15
Parra, M., Mora, A. et al. 2009a. Orogenic wedge Silver, P. G., Russo, R. M. & Lithgow-Bertelloni, C.
advance in the northern Andes: Evidence from the 1998. Coupling of South American and African
Oligocene-Miocene sedimentary record of the plate motion and plate deformation. Science, 279,
Medina Basin, Eastern Cordillera, Colombia. Geologi- 60–63.
cal Society of America Bulletin, 121, 780–800. Suppe, J., Chou, G. T. & Hook, S. C. 1991. Rates of
Parra, M., Mora, A., Sobel, E. R., Strecker, M. R. & folding and faulting determined from growth strata.
González, R. 2009b. Episodic orogenic-front In: McClay, K. R. (ed.) Thrust Tectonics, 1st edn.
migration in the northern Andes: constraints from low- Chapman & Hall, London, 105– 121.
temperature thermochronology in the Eastern Cordil- Taboada, A., Rivera, L. A. et al. 2000. Geodynamics of
lera, Colombia. Tectonics, 28, TC4004, http://dx.doi. the northern Andes: Subductions and intracontinental
org/10.1029/2008TC002423. deformation (Colombia). Tectonics, 19, 787–813.
Parra, M., Mora, A., Jaramillo, C., Torres, V., Zeilin- Tesón, E., Mora, A. & Silva, A. In press. Relationship of
ger, G. & Strecker, M. R. 2010a. Tectonic Mesozoic graben development, stress, shortening
Downloaded from http://sp.lyellcollection.org/ at Dalhousie University on November 13, 2013

A. MORA ET AL.

magnitude, and structural style in the Eastern Cordil- Velandia, F., Acosta, J., Terraza, R. & Villegas, H.
lera of the Colombian Andes. In: Nemčok, M., 2005. The curent tectonic motion of the northern
Mora, A. & Cosgrove, J. W. (eds) Thick-Skin- Andes along the Algeciras Fault System in SW Colom-
Dominated Orogens: From Initial Inversion to Full bia. Tectonophysics, 399, 313–329.
Accretion. Geological Society, London, Special Pub- Villamil, T. 1999. Campanian – Miocene tectonostrati-
lications, 377, http://dx.doi.org/10.1144/SP377.10 graphy, depocenter evolution and basin development
Uba, C. E., Kley, J., Strecker, M. R. & Schmitt, A. K. of Colombia and western Venezuela. Palaeogeogra-
2009. Unsteady evolution of the Bolivian subandean phy, Palaeoclimatology, 153, 239– 275.
thrust belt: the role of enhanced erosion and clastic Wetzel, A., Allenbach, R. & Allia, V. 2003. Reacti-
wedge progradation. Earth and Planetary Sciences vated basement structures affecting the sedimentary
Letters, 281, 134–146. facies in a tectonically ‘quiescent’ epicontinental
Van der Hammen, T., Werner, J. H. & Van Dommelen, basin: an example from NW Switzerland. Sedimentary
H. 1973. Palynological record of the upheaval of Geology, 157, 153–172.
the Northern Andes: a study of the Pliocene and Wijninga, V. M. 1996. Palynology and paleobotany of
Lower Quaternary of the Colombian Eastern Neogene sediments from the high plain of Bogotá
Cordillera and the early evolution of its High-Andean (Colombia): evolution of the Andean flora from an
biota. Reviews of Paleobotany and Palynology, 16, ecological perspective. PhD thesis, University of
1 –122. Amsterdam.

You might also like