You are on page 1of 29

Journal of South American Earth Sciences 21 (2006) 383–411

www.elsevier.com/locate/jsames

Mesozoic transtensional basin history of the Eastern Cordillera,


Colombian Andes: Inferences from tectonic models
a,*
L.F. Sarmiento-Rojas , J.D. Van Wess b, S. Cloetingh c

a
Ecopetrol-Empresa Colombiana de Petroleos, P.O. Box 5938-6813-Bogota, Colombia
b
Netherlands Institute of Applied Geoscience TNO National Geological Survey, Prins Hendriklaan 105 P.O. Box 80015 3508 TA Utretch, The Netherlands
c
Tectonics Group, Faculty of Earth Sciences, Free University, De Boelelaan 1085, 1081 HV Amsterdam, The Netherlands

Received 1 October 2004; accepted 1 March 2006

Abstract

Backstripping analysis and forward modeling of 162 stratigraphic columns and wells of the Eastern Cordillera (EC), Llanos, and
Magdalena Valley shows the Mesozoic Colombian Basin is marked by five lithosphere stretching pulses. Three stretching events are
suggested during the Triassic–Jurassic, but additional biostratigraphical data are needed to identify them precisely. The spatial distri-
bution of lithosphere stretching values suggests that small, narrow (<150 km), asymmetric graben basins were located on opposite
sides of the paleo-Magdalena–La Salina fault system, which probably was active as a master transtensional or strike-slip fault system.
Paleomagnetic data suggesting a significant (at least 10) northward translation of terranes west of the Bucaramanga fault during the
Early Jurassic, and the similarity between the early Mesozoic stratigraphy and tectonic setting of the Payandé terrane with the Late
Permian transtensional rift of the Eastern Cordillera of Peru and Bolivia indicate that the areas were adjacent in early Mesozoic times.
New geochronological, petrological, stratigraphic, and structural research is necessary to test this hypothesis, including additional
paleomagnetic investigations to determine the paleolatitudinal position of the Central Cordillera and adjacent tectonic terranes during
the Triassic–Jurassic. Two stretching events are suggested for the Cretaceous: Berriasian–Hauterivian (144–127 Ma) and Aptian–
Albian (121–102 Ma). During the Early Cretaceous, marine facies accumulated on an extensional basin system. Shallow-marine sed-
imentation ended at the end of the Cretaceous due to the accretion of oceanic terranes of the Western Cordillera. In Berriasian–Hau-
terivian subsidence curves, isopach maps and paleomagnetic data imply a (>180 km) wide, asymmetrical, transtensional half-rift basin
existed, divided by the Santander Floresta horst or high. The location of small mafic intrusions coincides with areas of thin crust
(crustal stretching factors >1.4) and maximum stretching of the subcrustal lithosphere. During the Aptian–early Albian, the basin
extended toward the south in the Upper Magdalena Valley. Differences between crustal and subcrustal stretching values suggest some
lowermost crustal decoupling between the crust and subcrustal lithosphere or that increased thermal thinning affected the mantle lith-
osphere. Late Cretaceous subsidence was mainly driven by lithospheric cooling, water loading, and horizontal compressional stresses
generated by collision of oceanic terranes in western Colombia. Triassic transtensional basins were narrow and increased in width
during the Triassic and Jurassic. Cretaceous transtensional basins were wider than Triassic–Jurassic basins. During the Mesozoic,
the strike-slip component gradually decreased at the expense of the increase of the extensional component, as suggested by paleomag-
netic data and lithosphere stretching values. During the Berriasian–Hauterivian, the eastern side of the extensional basin may have
developed by reactivation of an older Paleozoic rift system associated with the Guaicáramo fault system. The western side probably
developed through reactivation of an earlier normal fault system developed during Triassic–Jurassic transtension. Alternatively, the
eastern and western margins of the graben may have developed along older strike-slip faults, which were the boundaries of the accre-
tion of terranes west of the Guaicáramo fault during the Late Triassic and Jurassic. The increasing width of the graben system likely
was the result of progressive tensional reactivation of preexisting upper crustal weakness zones. Lateral changes in Mesozoic sediment
thickness suggest the reverse or thrust faults that now define the eastern and western borders of the EC were originally normal faults
with a strike-slip component that inverted during the Cenozoic Andean orogeny. Thus, the Guaicáramo, La Salina, Bitúima,
Magdalena, and Boyacá originally were transtensional faults. Their oblique orientation relative to the Mesozoic magmatic arc of

*
Corresponding author. Tel: +2345657.
E-mail address: luis.sarmiento@ecopetrol.com.co (L.F. Sarmiento-Rojas).

0895-9811/$ - see front matter  2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jsames.2006.07.003
384 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

the Central Cordillera may be the result of oblique slip extension during the Cretaceous or inherited from the pre-Mesozoic structural
grains. However, not all Mesozoic transtensional faults were inverted.
 2006 Elsevier Ltd. All rights reserved.

Keywords: Rifting; Lithosphere stretching; Tectonic subsidence; Colombia; Mesozoic; Eastern Cordillera

1. Introduction extensional episodes, plate-tectonic events, magmatic


events, and basin geometry. Many features of these exten-
This article focuses on the tectonic basin-forming pro- sional basins and their underlying mechanisms are practi-
cesses of the Colombian Eastern Cordillera (EC, Fig. 1) cally unknown.
during Mesozoic times, in terms of the geodynamic pro- Fabre (1983a,b, 1987) and Hébrard (1985) study the
cesses that govern deformation of the lithosphere. We com- subsidence of the eastern flank of the EC during the Creta-
pile local data into a regional geological model, analyze ceous, identify the basin as produced by lithosphere exten-
subsidence, and quantitatively model tectonic subsidence. sion, calculate tectonic subsidence curves, and, following
We have studied tectonic subsidence signals that give the uniform instantaneous stretching model developed by
important information about basin formation mechanisms. McKenzie (1978), calculate lithosphere stretching factors
For this purpose, we analyze temporal and spatial basin close to 2. They distinguish an Early Cretaceous subsidence
subsidence patterns, quantitatively analyze tectonic subsi- phase produced by rifting and Late Cretaceous subsidence
dence, and forward model it to explain these patterns from produced by thermal decay after rifting. We assume several
within the framework of the geodynamic processes that events of lithosphere stretching of finite duration in our
formed the Mesozoic EC basin. In doing so, we address study of tectonic subsidence and examine the possibility
issues such as the relationship among basin development, of differentiating between crustal and subcrustal stretching

Fig. 1. Location map. SL, Serranı́a de San Lucas; MA, Serranı́a de La Macarena. Eastern Cordillera regions: SM, Santander Massif; SF, Santander-
Floresta high; MT, Magdalena Tablazo inverted subbasin; CO, Cocuy inverted subbasin; MF, Magdalena Valley foothills; LF, Llanos foothills; CU,
Cundinamarca inverted subbasin; SE, Southern Eastern Cordillera; QM, Quetame Massif; GM, Garzón Massif; Romeral paleosuture in this map is the
western boundary of the Central cordillera and the Lower Magdalena Valley.
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 385

Fig. 2. Location of stratigraphic columns and wells and stratigraphic regional sections listed in Table 2. Numbers along sections refer to stratigraphic
transect labels (for details, see Sarmiento, 2001).

that occurred in the Colombian Basin throughout the basin existed, related to the topographic load of the Central
Mesozoic. An extensive data set of 162 stratigraphic col- Cordillera (e.g., Cooper et al., 1995). General consensus
umns and wells from the EC, Magdalena Valley (MV), indicates that during the Neogene, the Mesozoic extension-
and Llanos Orientales (LLA) areas (Fig. 2, see references al basin became inverted, deformed, and uplifted to form
in table 2.1 of Sarmiento, 2001) extracted from the litera- the EC (Cooper et al., 1995).
ture, as well as data from Ecopetrol, are used. In the study area, during the Triassic and Jurassic, con-
tinental and volcanic facies were deposited in extensional
2. Tectonic setting basins (Mojica et al., 1996). During the Triassic, these
basins seem related to Pangea rifting (Pindell and Dewey,
The EC is the eastern branch of the Colombian Andes 1982; Ross and Scotese, 1988; Cediel et al., 2003), and since
(Fig. 1), which comprises three mountain ranges: the East- the Jurassic, they developed behind a magmatic arc related
ern, Central, and Western Cordilleras, which merge south- to the subduction of the Pacific plates under the western
ward into a single range. The MV separates the Eastern border of South America (McCourt et al., 1984; Fabre,
and Central Cordilleras, and the Cauca Valley separates 1987; Toussaint and Restrepo, 1989; Cooper et al., 1995;
the Central and Western Cordilleras. The EC and its Meschede and Frisch, 1998). During the Early Cretaceous,
bounding basins, the LLA in the east and MV in the west, marine facies accumulated on a wide extensional basin sys-
define the area studied herein. tem. Shallow-marine sedimentation ended at the end of the
During the Mesozoic, the area of the EC was an exten- Cretaceous, due to the accretion of the oceanic terranes of
sional basin. During the Paleogene, according to some the Western Cordillera.
authors (e.g., Van der Hamen, 1961; Roeder and Chamber-
lain, 1995; Gómez et al., 1999; Sarmiento, 2001), upthru- 2.1. Triassic and Jurassic plate tectonic interpretations
sted blocks and/or incipient inversion of the Mesozoic
extensional basin may have occurred in the area of the The tectonic settings discussed subsequently assume
EC. However, another view posits a unique simple foreland that accretion of tectonic terranes east of the Romeral
386 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

paleosuture and west of the Llanos Orientales Basin 3. Intracontinental rifting related to the opening of the
(Fig. 1) occurred during the Paleozoic. Following this Caribbean. Some authors (e.g., Geotec, 1992; Cediel
assumption for evolution in Colombia, the plate tectonic et al., 2003) suggest a NW-SE graben developed in the
interpretations rely on two hypotheses. Although these northern part of the Central Cordillera during the Early
interpretations differ in plate tectonic setting, an alternative Cretaceous. A poorly defined Cretaceous magmatic arc
hypothesis suggests these settings formed at different times. in the Central Cordillera is difficult to explain with this
hypothesis.
1. Intracontinental rifting related to the breakup of Pangea
(Pindell and Dewey, 1982; Ross and Scotese, 1988; Ced- During the latest Cretaceous (post-Santonian), all plate
iel et al., 2003) occurred during Triassic and Early Juras- tectonic interpretations propose a convergent margin west
sic times. This hypothesis probably is more applicable to of Colombia. The Caribbean plate was moving northeast-
the northern part of Colombia and Venezuela and their ward relative to South America, while the Farallon plate
separation from North America. was subducting west of southern Colombia (Pindell and
2. Backarc extension occurred behind a subduction-related Erikson, 1993; Pindell and Tabbutt, 1995).
magmatic arc (Maze, 1984; McCourt et al., 1984; Pindell
and Erikson, 1993; Toussaint, 1995a,b; Pindell and Tab- 3. Stratigraphy
butt, 1995; Meschede and Frisch, 1998). According to
this hypothesis, the study area was located at the margin 3.1. Triassic and Jurassic synrift sedimentation
of the continent when active subduction of oceanic
Pacific plates was occurring. This interpretation explains The Triassic and Jurassic sedimentary record is present
the Triassic and Early Jurassic extensional basins in the in several isolated outcrops (Fig. 3). Continental deposits
study area as backarc basins. with redbeds and volcanic effusive and pyroclastic deposits
are dominant, though some marine facies appear locally.
However, a recent paleomagnetic investigation sug- Triassic and Jurassic rocks were deposited in extensional
gests that along-plate margin translations of terranes basins mainly located in the Upper Magdalena Valley
might have taken place during the early Mesozoic (Bay- (UMV), Serranı́a de San Lucas, and the western flank of
ona et al., 2005), as inferred by Toussaint and Restrepo the EC (Mojica et al., 1996).
(1994). Bayona et al. (2005), on the basis of paleomag- Fig. 4 shows a stratigraphic synthesis modified after
netic data, suggest a significant northward translation Mojica et al. (1996; Fig. 6). Triassic and Jurassic sedi-
(at least 10) of terranes west of the Bucaramanga fault mentary rocks formed a sequence bounded by unconfo-
(Bucaramanga area, Floresta Massif and Upper Magda- rmities. The lower contact is marked by an
lena Valley terranes) with respect to the craton during unconformity, which is dominantly angular. The upper
the Early Jurassic and no significant paleolatitude anom- contact is dominantly unconformable but locally con-
alies since then. Additional paleomagnetic data are need- formable. Jurassic deposits consisting of clastic facies
ed to test and quantify the magnitude of translation of deposited in dominantly continental environments are
tectonic terranes (Bayona, personal communication, widely distributed. In those sections where there are
2005). some marine facies, they are under- and overlain by con-
tinental clastic facies. The fine-grained muddy marine
2.2. Cretaceous plate tectonic interpretations facies record local marine incursions during the Late Tri-
assic–Early Jurassic. Volcaniclastic, pyroclastic, and vol-
For the Cretaceous, three alternative hypotheses suggest canic lavas are mainly restricted to the upper part of
the processes that might have operated at different times: the Upper Triassic to the lower part of the Middle Juras-
sic (Mojica et al., 1996, Fig. 6).
1. Backarc extension (McCourt et al., 1984; Fabre, 1987; Facies and thickness similarities related to geographic
Toussaint and Restrepo, 1989; Cooper et al., 1995; positioning suggest that Triassic–Jurassic sedimentation
Meschede and Frisch, 1998). Key evidence for this occurred in two separate basin compartments, each with
hypothesis is the existence of a subduction-related mag- its own subsidence history and sedimentary fill (Figs. 3
matic arc. and 4).
2. Passive margin (Pindell and Erikson, 1993; Pindell and
Tabbutt, 1995). The scarcity of magmatic rocks in the 1. Upper Magdalena, Cienaga de Morrocoyal, and Sierra
basin seems to support this hypothesis, but a poorly Nevada terrane (region A, Fig. 3). This section corre-
defined Cretaceous magmatic arc (i.e., San Diego, Alta- sponds with the western part of the Chibcha terrane,
vista, and Cambumbia stocks; Restrepo et al., 1991; as defined by Toussaint (1995a,b), or the Payandé, San
Toussaint and Restrepo, 1994) in the Central Cordillera Lucas, and Sierra Nevada terranes, as proposed by
is difficult to explain. Alternatively, the Central Cordille- Etayo-Serna et al. (1983). In this compartment, two
ra may have been located farther south of its present marine incursions are recognizable: in the Late Triassic
position during the Cretaceous. (Norian?–Rhetian, Chicalá member of the Saldaña
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 387

Fig. 3. Location of Triassic–Jurassic strata outcrops and stratigraphic sections. Labeling of stratigraphic sections according to Fig. 4. Stratigraphic section
E represents the Triassic–Jurassic sedimentary record of the eastern part of the Chibcha terrane according to Toussaint (1995b). Stratigraphic section W
represents the Triassic–Jurassic sedimentary record of the western part of the Chibcha terrane, equivalent to the Payandé, San Lucas, and Sierra Nevada
terranes of Etayo-Serna et al. (1983) (modified from Toussaint, 1995b). Inset: location of study area.

Formation, UMV) and in the Early Jurassic (Sinemurian, 2. Eastern Cordillera (region B, Fig. 3). This area corre-
Morrocoyal, and Los Indios formations; ages according sponds to the eastern side of the Chibcha terrane, as
to Mojica et al., 1996). Continental sedimentation fol- defined by Toussaint (1995a,b). Marine influence on
lowed shallow-marine deposition in this compartment. deposition is located within the Lower Jurassic (Monte-
Volcanic-related facies are volumetrically more impor- bel Formation; ages according to Mojica et al., 1996,
tant here than in the compartment formed by outcrops Fig. 4). This compartment is characterized by an absence
in the EC area (region B, Fig. 3). On the basis of paleo- of Triassic-dated rocks and thick deposition of siliciclas-
magnetic data from the UMV, Bayona et al. (2005) pro- tic deposits on continental extensional basins (e.g., Gir-
pose that the UMV and San Lucas (Payandé-San Lucas ón Formation). Bayona et al. (2005), on the basis of
terranes; Etayo-Serna et al., 1983) were located south of paleomagnetic data from the Bucaramanga and Floresta
the equator before the Middle Jurassic and north of it areas, suggest part of this terrane located west of the
during the late Jurassic–middle Cretaceous, with a Bucaramanga fault was located close to the equator dur-
northward movement of at least 10 relative to the South ing the early Middle Jurassic and moved northward
American craton. about 4 relative to the stable craton. According to their
388 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Fig. 4. Triassic–Jurassic stratigraphic sections. Location of transects in Fig. 3. Vertical axis represents geological time, and horizontal axis represents
present-day horizontal distance (km) without palinspastic restoration (modified from Mojica et al., 1996). Geyer (1982), however, suggests poorly
fossiliferous facies (El Sudán Formation) in the Cienaga de Morrocoyal area (north of Serranı́a de San Lucas) are Triassic and correlate with the Luisa
Formation of the Payandé area.

reconstruction, during the Late Triassic–Middle Juras- oscillating relative tectonoeustatic level. Subsidence was
sic, the UMV and Ciénaga de Morrocoyal (Payandé- rapid (Fabre, 1983a,b, 1987), but shallow-water sedimenta-
San Lucas terranes) ‘‘basin compartment’’ was located tion suggests deposition kept pace with it.
south of the EC compartment. The basin was a wide graben system oriented approxi-
mately NNE-SSW, divided into two subbasins (Tablazo
and Cocuy, Fig. 6) by the Santander-Floresta paleo-Mas-
3.2. Cretaceous sedimentation sif. To the north, these subbasins continued to the Machi-
ques trough in the Mérida Andes of Venezuela and
Most exposed rocks in the EC are Cretaceous in age. Uribante trough in Serranı́a de Perijá (Julivert, 1968; Fab-
Fig. 5 shows a traverse time-stratigraphic cross-section of re, 1985, 1987). To the south, these subbasins joined as the
the basin. Cundinamarca subbasin (Bürgl, 1961), where the thickness
Cretaceous rocks, including locally the uppermost of the Cretaceous sections reaches a maximum (Figs. 2 and
Jurassic and Paleocene deposits, form a megasequence 5). Fabre (1987) and Sarmiento (1989) suggest the Cundi-
bounded by regional unconformities that are at least local- namarca trough was limited to the south and north by
ly angular. On a broad scale, Cretaceous rocks represent a NW-SE transfer paleofaults (Fig. 6). The N-S lateral
major transgressive–regressive cycle with a maximum changes of thickness and an extensional relay system south
flooding surface close to the Cenomanian–Turonian of the Cundinamarca trough support the existence of NW-
boundary, corresponding to the maximum Mesozoic SE-striking transfer faults.
eustatic level (Fabre, 1985; Villamil, 1993; Fig. 5). Superim- On the basis of the presence of Lower Cretaceous sedi-
posed on this large-scale trend, several smaller transgres- mentary rocks in the northern part of the Central Cordille-
sive–regressive cycles are present, suggesting an ra, Geotec (1992) suggests the existence of a NW-SE
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 389

Fig. 5. Cretaceous and Tertiary stratigraphic section. Location in Fig. 2. Vertical axis represents geological time according to the scale of Gradstein and
Ogg (1996); horizontal axis represents present-day horizontal distance (km) without palinspastic restoration. This section has been constructed from
earlier versions by Fabre (1985, 1987) and Cooper et al. (1995), modified according to sequence stratigraphy interpretations (Pimpirev et al., 1992; Fajardo
et al., 1993; Villamil, 1993, 1994; Etayo-Serna, 1994; Ecopetrol, 1994; Rolón and Carrero, 1995; Villamil and Arango, 1998).

graben, connected with the Cundinamarca subbasin. Com- (Etayo-Serna and Laverde-Montaño, 1985). The amalgam-
piled available thicknesses of these sediments for similar ation of channel beds indicates that the sediment supply
chronostratigraphic intervals, however, are significantly exceeded the rate of basin subsidence at the Jurassic–Creta-
thinner than those of the EC (Figs. 5, 7b, and 8a). If such ceous boundary. Bürgl (1967) suggests that an initial mar-
a graben existed, in terms of subsidence, it was a minor fea- ine incursion in the Cundinamarca subbasin flooded a
ture compared with the grabens in the EC. These Lower continental area with a desert climate, which provided con-
Cretaceous marine sedimentary rocks in the Central Cor- ditions for evaporite formation during the early stages of
dillera are strongly deformed and associated with mafic marine transgression. McLaughlin (1972) cites paleonto-
volcanic rocks of ocean affinity, which collectively have logical evidence of a Berriasian–Valanginian age for some
been named the Quebradagrande Complex by Nivia et al. evaporite occurrences. During the Berriasian, the sea flood-
(1996) and Nivia and Gómez (2005). These authors inter- ed the basin from the northern part of the Central Cordil-
pret the rocks as originated within a marginal basin with lera toward the Cundinamarca subbasin (Etayo-Serna
oceanic crust formation developed along an older suture et al., 1976). Then, the sea advanced from the Cundinamar-
between different metamorphic terranes of the Central Cor- ca subbasin to the north into two subbasins, while the
dillera (between the eastern Cajamarca Complex and the Santander-Floresta paleo-Massif remained emergent
western Arquı́a Complex). Compiled thicknesses suggest (Etayo-Serna et al., 1976; Fabre, 1985, 1987; Sarmiento,
a current N-S basin in agreement with Nivia and Gómez’s 1989; Figs. 5, 7a, b, and 8a).
(2005) interpretation. The Tablazo and Cocuy subbasins started to form a sin-
gle wide basin during the Hauterivian due to flooding of
3.2.1. Early cretaceous synrift sedimentation the Santander-Floresta paleohigh (Fabre, 1985) and base
Sedimentation started in the Tablazo subbasin in Juras- level rise. However, this intrabasinal high was a significant
sic times and continued during the Early Cretaceous local- barrier to sediment movement until the Aptian (Figs. 7a, b,
ly, without a tectonic-related angular unconformity (e.g., and 8a) .
Rio Lebrija section, Cediel, 1968). In other areas, Creta- To the south, both the Tablazo and Cocuy subbasins
ceous sedimentary rocks rest with angular unconformity show a gradual increase in dark shale deposited in poorly
on earlier Mesozoic, Paleozoic, or even pre-Cambrian oxygenated shallow-marine environments (Caqueza and
rocks. In the Tablazo subbasin, basal beds correspond to Villeta groups; Fabre, 1985; Sarmiento, 1989). In the Cun-
sandstones (Los Santos, Tambor, and Arcabuco dinamarca subbasin, Cretaceous sedimentation started
formations) deposited in fluvial environments, usually with during the Tithonian?–Berriasian–Valanginian with turbi-
detrital sediments derived from uplifted fault blocks dite deposits in both the eastern (lower Caqueza Group;
390 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Fig. 6. Cretaceous basin compartments (subbasins) and their tectonic subsidence in meters. (A) Tectonic subsidence of the Tablazo and Cocuy subbasins.
(B) Tectonic subsidence of the Cundinamarca subbasin. Cities: M, Medellı́n; Mz, Manizales; I, Ibagué; N, Neiva; C, Cúcuta; Bu, Bucaramanga; T, Tunja;
B, Bogotá; V, Villavicencio; Y, Yopal; A, Arauca.

Pimpirev et al., 1992) and western (lower part of Utica (Etayo-Serna et al., 1976; Etayo-Serna, 1994). Southward
sandstone, Murca Formation; Sarmiento, 1989; Moreno, of the paleo-UVM, Cretaceous sedimentation started dur-
1990, 1991) flanks (Figs. 7a, b, and 8a). Turbidite deposi- ing the Aptian (Vergara and Prössl, 1994) in an extensional
tion prevailed until the Hauterivian at the eastern border basin formed initially in Jurassic times.
of the basin (Caqueza Group, Pimpirev et al., 1992). In the whole EC basin, abrupt lateral thickness changes
During the earliest Cretaceous, basin subsidence and fault-related alluvial or turbidite deposition attest to
exceeded sediment supply, resulting in retrogradation of local tectonic/differential subsidence depositional condi-
the turbidite fan system, so distal fan sediments covered tions in the Early Cretaceous. Regional correlation of Ear-
middle fan mouth channel deposits. Post-Berriasian, sed- ly Cretaceous relative tectonoeustatic cycles is difficult to
iment supply increased and overwhelmed basin subsi- establish because of local active extensional tectonics. Since
dence, resulting in progradation of the turbidite fan the Aptian, these relative tectonoeustatic cycles have
system (Pimpirev et al., 1992) and locally in progradation become more tractable than the pre-Aptian cycles (Fig. 5).
of deltaic sands during the Hauterivian (upper Utica In the Central Cordillera, Lower Cretaceous sedimenta-
sandstone; Sarmiento, 1989; Moreno, 1990). Differential ry rocks (mudstones, chert, feldspathic sandstones, and
subsidence related to syn-sedimentary normal faulting conglomerates) of the Quebradagrande Complex are asso-
caused unstable slopes on basin margins. These processes ciated with an ophiolite suite of ultramafic rocks, gabbros,
favored turbidite deposition during the early Cretaceous– basalts, breccias, and pyroclastic rocks usually affected by
Aptian (lower Utica sandstone, Murca Formation, Sar- dynamic metamorphism (Moreno and Pardo, 2003). Depo-
miento, 1989; Moreno, 1990, 1991; Socota Formation, sitional environments vary among fluvial, coastal, deltaic,
Polanı́a and Rodrı́guez, 1978; Caqueza Group, Pimpirev and marine shelf (Rodrı́guez and Rojas, 1985). Evidence
et al., 1992; Figs. 7a, b, and 8a). indicates that during the middle Cretaceous, these rocks
An important transgression followed a relative sea-level began compressive deformation (Rodrı́guez and Rojas,
rise during late Aptian time. The sea flooded all the area of 1985), possibly as a result of the closure of the Quebrada-
the present EC, even south of the Cundinamarca subbasin grande marginal basin (Nivia et al., 1996).
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 391

Fig. 7. (a) Berriasian–Valanginian and (b) Hauterivian–Barremian thickness (m) without palinspastic restoration. Thick lines represent paleofaults
believed to be active during Berriasian–Valanginian/Hauterivian–Barremian, respectively. (c) Contour map of crustal (d) lithosphere stretching factors
calculated through forward modeling for the Berriasian–Hauterivian (144–127 Ma, Cretaceous) stretching event without palinspastic restoration.
Distribution of main Early Cretaceous faults and mafic intrusions shown with circles: 1. Rio Nuevo diorite, 2. Rodrigoque micrograbbro, 3. Porfiritic
basaltic lava, 4. Rio Cravo Sur microgabbro, 5. Pajarito, 6. Q. La Esperanza, 7. Q. Las Palomas, 8. Q. La Culebra, 9. Marfil, 10. Q. Grande, 11. Q. La
Chorrera, 12. La Chunchalita, 13. Q. La Fiebre, 14. Caceres, 15. La Corona, 16. Pacho, and 17. Rio Guacavia diorite. (d) Contour map of subcrustal (b)
lithosphere stretching factors calculated through forward modeling for the Berriasian–Hauterivian (144–127 Ma, Cretaceous) stretching event without
palinspastic restoration. Distribution of main Early Cretaceous faults is shown. Cities: M, Medellı́n; Mz, Manizales; I, Ibagué; N, Neiva; C, Cúcuta; Bu,
Bucaramanga; T, Tunja; B, Bogotá; V, Villavicencio; Y, Yopal; A, Arauca.

3.2.2. Cretaceous postrift sedimentation marine facies (e.g., Caballos, San Gil Inferior, and Socotá
Several relative tectoeustatic level cycles have been pro- formations) to outer shelf facies (e.g., Villeta Group, San
posed during the Late Cretaceous (Fig. 5). Gil Superior, and Hiló formations) recorded a rise in
During the Albian, a relative base-level fall favored pro- relative tectonoeustatic levels (Villamil, 1993; Etayo-Serna,
gradation of deltaic and littoral sands in the UMV (Cabal- 1994). During late Albian–early Cenomanian, a relative
los Formation, Etayo-Serna, 1994) and the eastern border tectonoeustatic level fall was recorded by progradation of
of the basin (lower Une Formation, Fabre, 1985). During the upper part of Une Formation and a generalized shal-
the middle–late Albian, the transition from near-shore lowing-upward facies trend (e.g., transition of San Gil
392 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Fig. 8. (a) Aptian thickness (m) without palinspastic restoration. Thick lines represent paleofaults believed to be active during Aptian time. (b) Contour
map of crustal (d) lithosphere stretching factors calculated through forward modeling for the Aptian (121–102.6 Ma, Cretaceous) stretching event without
palinspastic restoration. Distribution of main Early Cretaceous faults is also shown. (c) Contour map of subcrustal (b) lithosphere stretching factors
calculated through forward modeling for the Aptian (121–102.6 Ma, Cretaceous) stretching event without palinspastic restoration. Distribution of main
Early Cretaceous faults also shown. Cities: M, Medellı́n; Mz, Manizales; I, Ibagué; N, Neiva; C, Cúcuta; Bu, Bucaramanga; T, Tunja; B, Bogotá; V,
Villavicencio; Y, Yopal; A, Arauca.

Superior to Churuvita and Hiló to unnamed shale; Villa- the Campanian (CS at the top of Gachetá Formation; Faj-
mil, 1993). ardo et al., 1993; Cooper et al., 1995, Fig. 4).
During the late Cenomanian, Turonian, and Coniacian, Villamil (1993) recognized smaller relative tectonoeustat-
the tectonoeustatic base level reached its maximum Meso- ic level cycles during late Cenomanian, Turonian, and
zoic level. The sea flooded the entire northwestern corner Coniacian times. A relative tectonoeustatic base level rise
of South America, and dark gray shale was deposited from during the late Cenomanian (Villamil, 1993) induced a slight
Venezuela to northern Peru (Thery, 1982, in Fabre, 1985). deepening of the basin and a notorious decrease of detrital
In contrast to the EC, where the Cretaceous maximum supply to the basin (e.g., Frontera and lower part of San
flooding surface occurred at the Cenomanian–Turonian Rafael formations; Villamil, 1993). During the Turonian–
boundary, maximum flooding in the LLA occurred during Coniacian, the present-day LLA foothills flooded (Cooper
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 393

et al., 1995) but not the entire LLA (Fig. 5). From the Middle Table 1
Turonian to late Coniacian, a gradual progradation and Parameters used to calculate tectonic subsidence in the forward models
shallowing upward (e.g., upper San Rafael Formation and Model parameters Value
Villeta Group in the upper MV) was related to a relative tec- Initial lithospheric thickness 120 km
tonoeustatic level fall (Villamil, 1993). Initial crustal thickness 35 km
In the UMV and during the late Coniacian–Santonian, Asthenospheric temperature 1333 C
Thermal diffusivity 1 · 106 m2 s1
deepening of the basin and relative tectonoeustatic level Surface crustal density 2800 kg m3
rise occurred (transition from inner shelf uppermost Villeta Surface mantle density 3400 kg m3
Group to middle shelf lower chert unit of the Olini Group; Seawater density 1030 kg m3
Etayo-Serna, 1994, Fig. 2). Thermal expansion coefficient 3.2 · 105 C
During the Santonian, Campanian, Maastrichtian, and
Paleocene, a general regression and progradation was column with available geological maps to avoid structural
recorded by littoral to transitional coastal plain facies repetitions. We also verified the consistency of thickness
(e.g., Guadalupe Group, Guaduas Formation). Guadalupe between neighboring sections. The effects of paleobathyme-
Group sands represent two cycles of westward shoreline try have been taken into account, using sedimentary facies
progradation, aggradation, and retrogradation, dominated and faunal content as interpreted in the literature. Ages are
by high-energy, quartz-rich shoreface sandstones derived based on data in the literature, mainly the regional strati-
from the Guyana Shield (Cooper et al., 1995; Fig. 5). graphic cross-sections presented by Cooper et al. (1995).
Regression did not occur continuously but with minor To express ages in Ma, we used the geological time scale
transgressive events recorded by fine-grained siliceous and proposed by Gradstein and Ogg (1996). Unconformities
phosphatic facies (Plaeners Formation, Olini Group, and are also included with ages and durations in Ma. Addition-
upper La Luna Formation, Föllmi et al., 1992; Fig. 5). al parameters for forward modeling are the initial crust and
At the end of Cretaceous (Maastrichtian), the gradual lithosphere thickness and densities. Table 1 provides the
uplift of the western margin of the MV supplied clasts of additional parameters we used in forward modeling, which
metamorphic rocks accumulated by fluvial systems close are the average accepted values for normal continental lith-
to the sea in a braided delta (Cimarrona Formation, osphere (Table 2).
Gómez and Pedraza, 1994). Pulses of fast tectonic subsidence of the basement have
In the Central Cordillera east of the Quebradagrande been interpreted in terms of tectonic processes. Only those
Complex, no Late Cretaceous sedimentary record has been pulses of fast tectonic subsidence correlatable to the nor-
found. This area may have been uplifted during the latest mal movement component of fault activity are interpreted
Cretaceous and started to supply clastics to the east. as produced by lithosphere extension or transtension,
though normal faulting also may correspond to transte-
4. Methods sional settings. Fast subsidence events related to crustal
or lithosphere thinning also is common in transtensional
4.1. Subsidence analysis basins or intra-arc basins (Ingersoll and Busby, 1995). In
these settings, the extensional component is a major con-
The stratigraphic record provides information about tributor to tectonic subsidence. The slower, generally later,
vertical crustal movements in a basin. Basin subsidence is rate of subsidence has been interpreted as produced by
the result of both a thermomechanical component called thermal reequilibration of the lithosphere following the
tectonic subsidence and a sediment and water loading com- thermal anomaly created by stretching.
ponent. Tectonic subsidence is undistorted basin subsi-
dence in the absence of sedimentation and therefore is 4.2. Forward modeling of basin evolution
related to the geodynamics of the basin.
To quantify the tectonic component of subsidence of the To quantify the horizontal extensional movements
studied basin, we used the one-dimensional (1D) backstrip- responsible for the observed subsidence and establish a
ping technique (Steckler and Watts, 1978; Bond and quantitative framework for the pulsating rift evolution of
Kominz, 1984), as explained by Sclater and Christie (1980), the lithosphere during Mesozoic basin formation, we quan-
Bond and Kominz (1984), and Bessis (1986). For this pur- tify extension rates by forward modeling tectonic subsi-
pose, we calculate tectonic subsidence from the stratigraphic dence. We use an ‘‘automated’’ forward modeling
record, adopting local Airy isostasy to correct for the effect of technique (Van Wees et al., 1996b), which we explain brief-
sediment loading. The corrections for compaction follow ly next.
porosity–depth relationships on the basis of the observed
lithologies using standard mean exponential relations and 4.2.1. Numerical model
material parameters (cf. Sclater and Christie, 1980). The forward modeling approach is based on lithospheric
Most stratigraphic columns are from published litera- stretching assumptions (McKenzie, 1978; Royden and
ture (see Sarmiento, 2001); well data are from Ecopetrol. Keen, 1980). The extension factor d is used for crustal
We carefully checked the thickness of each stratigraphic stretching and b for subcrustal stretching. For thermal
394 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Table 2
Cretaceous stretching events and stretching lithosphere factors from stratigraphic columns where the Cretaceous sedimentary record is present
Basin compartmenta Stratigraphic columnb Berriasian–Hauterivian Event Aptian–early Albian event
Start (Ma) End (Ma) Stretching Start (Ma) End (Ma) Stretching
factor factor
b d b d
2 72 Casabe-199 144 127.8 1.522 1.146 114 109.3 1.03 1.03
c
2 74 Cascajales-1 144 127.5 0.987 0.987 114 109.3 1.132 1.094
2 73 Infantas-1613 138 127.8 1.301 1.117 114 109.3 1.328 1.113
2 53 Lebrija-1 144 127 1.625 1.213 114 109.3 1 1
2 71 Llanito-1 138 127.8 1.313 1.098 114 109.3 1.478 1.131
2 70 Sabalo-1K 144 127.8 1.37 1.113 114 109.3 1.039 1.039
c
3 4 Arcabuco 140 127.5 1.099 1.099
c
3 8 Chima 144 127.5 1.47 1.261 114.8 112.2 1.063 1.019
3 10 Cimitarrra 143 127.5 1.354 1.31 114.8 112.2 1.329 1.111
c
3 24 Los Medios 142 127.5 1.043 1.043 114.8 112.2 1.096 1.043
3 25 Los Santos 136.5 127.5 1.131 1.088 114.8 112.2 1.009 1.009
3 47 Simacota 144 127.5 1.251 1.251 114.8 112.2 1.064 1.01
c
3 52 Tablazo 144 127.5 1.052 1.052 114.8 112.2 1 1
3 57 Vadorreal 134.5 127.5 1.585 1.179 114.8 112.2 1.374 1.08
3 58 Velez 142 127.5 1.151 1.245 114.8 112.2 1.219 1.072
3 59 Villa de Leiva 142 127.5 2.016 1.332 114.8 112.2 1.052 1.052
4 66 Chitasuga-1 142 130 1.196 1.196 121 112.2 1.536 1.187
4 22 La Calera 142 130 1.331 1.206 121 112.2 1.58 1.178
c
4 39 Quipile 141 130 1.511 1.147 121 112.2 1.301 1.115
4 48 Simijaca 142 130 1.247 1.247 121 112.2 3.238 1.318
c
4 75 Suba-2 142 130 1.085 1.085 121 112.2 1.221 1.142
4 76 Suesca-1 142 130 1.217 1.217 121 112.2 2.476 1.304
4 77 Suesca Norte-1 142 130 1.214 1.214 121 112.2 1.824 1.224
4 50 Sutamarchan 142 130 1.234 1.234 121 112.2 1.825 1.252
4 51 Tabio 142 130 1.227 1.227 121 112.2 1.587 1.206
c
4 60 Villeta 141 127.5 1.01 1.01 121 112.2 1.52 1.492
c
4 61 Yacop 134.2 127.5 1.449 1.449 121 112.2 1.77 1.299
5 3 Apulo 127.6 127 1.792 1.164 121 112.2 1.261 1.049
5 15 Fusagasuga 127.6 127 1.571 1.13 121 112.2 1.466 1.071
c
6 2 Alpujarra 115.1 102.6 1.411 1.222
c
6 12 Coello 121 102.6 1.11 1.066
6 16 Girardot 121 102.6 1.366 1.157
6 19 Guataqui 121 102.6 1.366 1.173
6 21 Itaibe 108.5 102.6 1.176 1.087
6 17 Melgar 121 102.6 1.18 1.121
c
6 29 Neiva 121 102.6 1.312 1.082
6 31 Ortega 121 102.6 1.658 1.284
c
6 35 Prado 121 102.6 1.422 1.115
6 36 Q Calambe 119.1 102.6 1.477 1.1
c
6 37 Q El Cobre 121 102.6 1.275 1.024
c
6 38 Q Olini 121 102.6 1.166 1.049
7 56 Chivata 132.8 127 3.144 1.215
7 64 Cormichoque-1 122.4 122.3 2.029 1.179
7 14 Floresta 132.6 127 1.024 1.037
7 18 Guaca 132 127 1.259 1.075
7 26 Matanza 128.8 127 1.716 1.13
7 54 Tibasosa 132.8 127 2.046 1.181
7 55 Tunja 133.5 127 2.204 1.185
7 78 Tunja-1 132 127 2.576 1.213
8 63 Bolivar-1 Corrales-1 132 127 2.071 1.186
c
8 6 Caqueza 144 127.5 1.215 1.204
8 28 Nazareth 131.5 127.5 1.493 1.155
8 34 Paz de Rio 144 127 1.385 1.154
8 44 Servita 132 127.5 1.565 1.138
9 33 Aguazul 142 127.5 1.205 1.309
9 9 Chita 139 127.5 2.525 1.482
9 7 Cocuy 139 127.5 3.489 1.657
9 20 Guateque 138 127.5 1.035 1.405
9 23 Labateca 132 127.5 1.478 1.285
9 69 Medina-1 144 127.5 3.605 1.118
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 395

Table 2 (continued)
Basin compartmenta Stratigraphic columnb Berriasian–Hauterivian Event Aptian–early Albian event
Start (Ma) End (Ma) Stretching Start (Ma) End (Ma) Stretching
factor factor
b d b d
9 27 Mojicones 139 127.5 1.36 1.23
9 32 Pajarito 142 127.5 1.169 1.303
9 40 R Cusay 136.5 127.5 1.643 1.35
c
9 46 San Luis de Gaceno 144 127.5 1.076 1.352
9 49 Sogamoso 136.5 127.5 1.459 1.354
10 67 Cusiana-1X-2 112.3 112.2 1.095 1.025
Basin compartment numbers as shown in Fig. 6. Location of each stratigraphic column identified by number shown in Fig. 2.
a
Numbers indicate Cretaceous subbasins shown in Fig. 6.
b
Stratigraphic column number shown in Fig. 2.
c
Modeled using Triassic and Jurassic actual or inferred stratigraphy.

Fig. 9. Outline of the forward modeling technique. Explanation in the text (from Van Wees et al., 1996b).

calculations, we use a 1D numerical finite difference model, two-layered stretching (d „ b) (e.g., Royden and Keen,
which allows us to incorporate finite and multiple stretch- 1980) can be used. For uniform stretching, the solution
ing phases. To handle the large number of wells and stretch- to Eq. (1) requires that at least one observed subsidence
ing phases in the forward model, we apply a numerical data point is given after the onset of rifting, whereas two-
technique (Van Wees et al., 1996b), which automatically layered stretching requires at least two data points. For
finds the best fit stretching parameters for (part of) the sub- polyphase stretching, the fit is accomplished in sequential
sidence data. In this procedure, we must specify the timing order. Initially, using a steady state thermal and composi-
and duration of the rift phase, whereas the best fit stretch- tional lithospheric configuration (cf. McKenzie, 1978),
ing values emerge by searching for the minimum of the stretching parameters of the first phase are determined by
mean square root F of the deviation in predicted and fitting data points in the syn- and postrift time interval to
observed subsidence (Fig. 9), as a function of d and b, as the onset of the following phase. Subsequently, using the
follows: perturbed lithosphere configuration predicted at the onset
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi of the second rift phase, stretching parameters are deter-
uinum
1 u X
F ðd; bÞ ¼ t ðspp;i  so;i Þ ;
2
ð1Þ mined using subsidence data from its syn- and postrift time
num i1 intervals to the next rifting phase.

where num is the number of subsidence data used in the fit- 4.2.2. Modeling procedure
ting procedure, and sp,i and so,i are predicted and observed In the fitting procedure, initial lithospheric configuration
subsidence values, respectively. For a rift phase, either uni- and thermal parameters are adopted as listed in Table 1.
form lithospheric stretching (d = b) (McKenzie, 1978) or To fit the data, we assume that each observed phase of
396 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

rapid tectonic subsidence should correspond to a stretching 6.2. Triassic–Jurassic subsidence and lithosphere stretching
phase in the forward model. For these phases, we adopt a events
two-layered stretching model of the lithosphere (d „ b) to
obtain the highest degree of freedom. However, we prefer 6.2.1. Early? Triassic event (variable in different columns,
to use a uniform stretching model (d = b) when uncertainty 248 to 235 Ma, time scale of Gradstein and Ogg, 1996)
exists in estimating the stretching factors due to fewer data This event is best represented in the UMV–Ciénaga de
points or for relatively large age uncertainties, as is the case Morrocoyal with uniform stretching factors; b = d reach
for the Triassic and Jurassic sedimentary record. values of 1.23. Subsidence curves (Fig. 10), thickness
Using the starting and finishing times previously deter- variations, and the spatial distribution of stretching val-
mined for the stretching events, we calculate the litho- ues (Fig. 12a) and paleomagnetic data (Bayona et al.,
sphere stretching factors that would produce theoretical 2005) suggest that small, narrow (150 km), transcurrent
subsidence curves similar to those observed. For the for- rift basins formed. Abrupt lateral changes of sediment
ward modeling, we include for most of the modeled loca- thickness and facies in the UMV suggest differential sub-
tions the complete Mesozoic sedimentary section since sidence in different faulted blocks (Bayona et al., 1994;
the Triassic, even in those columns in which the pre-Me- Mojica et al., 1996), which is common in transcurrent
sozoic section is probably deep and does not crop out. In faults. Jurassic normal faults are described by Guillande
these cases, we use thicknesses interpolated from isopach (1988).
maps. The transtensional event correlates in time with mag-
matic arc activity in the Central Cordillera, interpreted
5. Results by Aspden et al. (1987) as related to oblique subduction.
The event also may have been related to intracontinental
In Fig. 10, we show the tectonic subsidence curves. The rifting (breakup of Pangea), as proposed elsewhere, par-
forward-modeled tectonic subsidence curves (Fig. 11) indi- ticularly the separation between South and North Amer-
cate a remarkably good fit with the subsidence data. The ica (e.g., Pindell and Dewey, 1982; Ross and Scotese,
lithosphere, crustal, and subcrustal stretching factors calcu- 1988; Cediel et al., 2003). However, considering that rif-
lated for each stretching phase also are plotted in map view ting and subsequent sea floor spreading, which originated
(Figs. 7, 8, and 12). the Gulf of Mexico, occurred during the Late Jurassic
(Pessagno and Martin, 2003), this event probably relates
6. Tectonic subsidence and lithosphere stretching during to local transtensional events along the continental
Triassic and Jurassic time margin.

For the tectonic subsidence of the Triassic sedimenta- 6.2.2. Latest Triassic–Middle Jurassic event (208 to
ry record, we assume, following Geyer (1982), that the 185 Ma)
poorly fossiliferous sediments (El Sudán Formation) of Subsidence curves (Fig. 10), stratigraphic thickness, map
the Cienaga de Morrocoyal – north of the Serranı́a de distribution of stretching values, paleomagnetic data (Bay-
San Lucas with lithology and relative stratigraphical ona et al., 2005), and fault distribution (Fig. 12b) suggest
position similar to the Luisa Formation of the Payandé two narrow (<150 km), transtensional basins located at
region – are Triassic and time correlative and that Trias- the current location of Serranı́a de San Lucas (Fig. 1)
sic sediments accumulated in the western flank of the and the UMV. Fast subsidence favored marine ingression
EC. Using subsidence patterns, we attempt to differenti- mainly in the south (Payandé terrane). Probably in a dom-
ate fast subsidence events. However, these attempts inant strike-slip setting, transtensional basins were separat-
should be considered preliminary, because the continental ed by transpressional nodes, depending on the geometry of
Triassic and Jurassic poorly fossiliferous sedimentary the strike-slip faults.
record has relatively scarce and low-quality biostrati- The abundance of volcanic rocks in these transtensional
graphic data that make it difficult to establish clear time basins implies a positive thermal anomaly that probably
boundaries between the events. weakened the lithosphere. Lithosphere thermal doming
could have produced the observed unconformities at the
6.1. Basin compartments bottom of the synrift fill. Fast subsidence rates and a high
level of volcanic activity are features commonly associated
Differences in the shape of the subsidence curves in dif- with oblique slip-rift zones (Ziegler, 1994). The width of
ferent areas confirm that Triassic–Jurassic sedimentation these basins increased compared with the early Triassic
occurred in two separate basin compartments, each with (Mojica et al., 1996), suggesting the increasing width of
its own subsidence history and sedimentary fill: the thermal weakened lithosphere. A weak lithosphere
favored transtensional deformation.
1. Payandé-San Lucas terranes (UMV and Ciénaga de This extensional event correlates with magmatic arc
Morrocoyal, region A, Fig. 3). activity in the Central Cordillera and may be interpreted
2. Eastern Cordillera (region B, Fig. 3). as a backarc extension (Aspden et al., 1987). Rifting
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 397

and subsequent sea floor spreading, which originated the 6.2.3. Middle Jurassic event (180 to 176 Ma)
Gulf of Mexico, occurred during the Late Jurassic Paleogeography and stratigraphic thickness distribu-
(Pessagno and Martin, 2003), so this event probably is tions indicate continued widening of transtensional basins,
related to local transtensional events along the continen- though they remained relatively narrow. Major depocen-
tal margin. ters developed in the MV and western flank of EC within

Fig. 10. Tectonic subsidence curves from the Mesozoic sedimentary record. Horizontal axis represents age in Ma. Vertical axis represents tectonic
subsidence in meters obtained from backstripping analysis. Vertical shaded strips represent fast subsidence events. Numbers refer to basin compartments
in Fig. 6. Note vertical bars represent the fast tectonic subsidence events.
398 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Fig. 10 (continued)

elongated NNE transcurrent grabens on opposite sides of netic data (Bayona et al., 2005) indicate rift basins located
the Magdalena–La Salina fault system. The distribution along the present-day western flank of the EC, with b = d
of stretching b = d factor values (Fig. 12c) and paleomag- values up to 1.39, and the paleo-MV. Moreover, large
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 399

Fig. 11. Forward modeled tectonic subsidence (continuous line) and observed tectonic subsidence (dots) curves (in m).

postrift Cretaceous subsidence along the western flank of However, considering that rifting and subsequent sea floor
the EC northwest of Bogotá can be explained only as ther- spreading, which originated the Gulf of Mexico, occurred
mal subsidence after a Jurassic stretching event. Therefore, during the Late Jurassic (Pessagno and Martin, 2003), these
these basins extended southward into the Cundinamarca events probably relate to local transtensional events along
region, possibly limited by the paleo-La Salina, paleo- the continental margin. Paleomagnetic data suggest that
Suarez, and paleo-Boyaca fault systems. Small, isolated during the Late Triassic and Early Jurassic, the Santa Marta,
grabens developed in the Santander Massif (Geotec, Santander Massif, Floresta Massif, San Lucas, and Payandé
1992; Kammer, 1993), Perijá (Shagam, 1975; Maze, (UMV) magmatic arc segments aligned into a magmatic arc,
1984), Mérida Andes (Ricardi et al., 1990, in Mojica striking parallel to the subduction zone (Bayona et al., 2005).
et al., 1996), LLA (Numpaque, 1986, in Cooper et al., This magmatic arc belt formed as a result of subduction
1995; Geotec, 1992), and Maracaibo (Shubert and Ricardi, along western Pangea, which resulted in the southward con-
1980, in Mojica et al., 1996) areas. Volcanic activity tinuation of the early Mesozoic continental magmatic arc
decreased at this time, mainly occurring in the Mérida from the southwestern United States to Guatemala (Bayona
Andes (basalts in La Quinta Formation, Maze, 1984). et al., 2005). In the eastern part of the Chibcha terrane (EC
These Triassic–Jurassic events correlate with magmatic east of the Bucaramanga fault), the main calc-alkaline plu-
arc activity in the Central Cordillera and Santander Massif. tonic bodies developed on the Santander Massif are posi-
The calc-alkaline composition of these magmatic rocks sug- tioned between the transtensional basins, suggesting an
gests a magmatic arc related to high-angle subduction (Asp- interarc setting.
den et al., 1987). The most developed calc-alkaline magmatic
plutonic bodies of the Payandé-San Lucas terranes (i.e., 7. Tectonic subsidence and lithosphere stretching during
Ibagué and Segovia batholiths) tend to be adjacent to the latest Jurassic and Cretaceous
west of these transtensional basins, which indicates these
transtensional basins were located in a backarc setting close 7.1. Basin compartments
to the arc or an intraarc setting. Alternatively, these events
may have been related to intracontinental rifting (breakup Tectonic subsidence curves (Fig. 10) and restored thick-
of Pangea), as proposed elsewhere, particularly the separa- ness maps (Figs. 7a, b, and 8a) indicate several basin com-
tion between South and North America (e.g., Pindell and partments (Fig. 6). In the northern part of the EC (Fig. 6),
Dewey, 1982; Ross and Scotese, 1988; Cediel et al., 2003). we recognize
400 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Fig. 12. (a) Contour map of lithosphere stretching factors (b = d) calculated through forward modeling for the Triassic (248.2–235 Ma) stretching event,
assuming there are Triassic sediments in the Cienaga de Morrocoyal area (Geyer, 1982) and that Triassic sediments accumulated in the western flank of the
EC, without palinspastic restoration. Distribution of main early Mesozoic faults also shown. (b) Contour map of lithosphere stretching factors (b = d)
calculated through forward modeling for the Early Jurassic (208–185 Ma) stretching event without palinspastic restoration. Distribution of main early
Mesozoic faults is shown. (c) Contour map of lithosphere stretching factors (b = d) calculated through forward modeling for the Jurassic (180.1–176 Ma)
stretching event without palinspastic restoration. Distribution of main Early Mesozoic faults also shown. (A) Payandé, San Lucas terranes (Etayo-Serna
et al., 1976), western part of Chibcha terrane (Toussaint, 1995a); (B) eastern part of Chibcha terrane (Toussaint, 1995a) and Guyana shield. Cities: M,
Medellı́n; Mz, Manizales; I, Ibagué; N, Neiva; C, Cúcuta; Bu, Bucaramanga; T, Tunja; B, Bogotá; V, Villavicencio; Y, Yopal; A, Arauca.

1. Two subbasins: the Cocuy (region 9) and In the southern part of the Cordillera at the latitude of
Tablazo (region 3) graben subbasins, separated Bogotá, we recognize
by the less subsiding Santander-Floresta block
(region 7), and
2. A regional westward decrease in tectonic subsidence 1. A single extensional basin, the Cundinamarca subbasin
with a maximum in the Cocuy subbasin (region 9) (region 4, Fig. 6), and
and a minimum in the Middle MV (region 2), which 2. Berriasian–Hauterivian tectonic subsidence, maximal in
suggests a regional half-graben geometry for the whole the eastern side of the Cundinamarca subbasin (region
basin. 8, Fig. 6), which indicates that a first stretching event
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 401

mainly affected the eastern Guaicáramo normal fault normal faulting. Normal faults imaged on seismic sections
system. During the Aptian, however, subsidence was (Fig. 13) confirm extensional tectonic movements that attest
maximal in the western side (region 4, Fig. 6), so a sec- that this rapid subsidence event was produced by litho-
ond stretching event mainly affected the western Bituima sphere stretching. As suggested by paleomagnetic data
fault system. The total tectonic subsidence during the (Bayona et al., 2005), these faults may have had a dextral
Cretaceous was slightly greater on the western side of strike-slip component. Unlike the Triassic–Jurassic trans-
the basin (region 4). tension, magmatic activity within the basin was reduced
during the Early Cretaceous. Evidence of Early Cretaceous
In the western part on the Central Cordillera, a N-S magmatism is limited to small, mafic, igneous intrusions
(present day geometry) depocenter represents the Quebra- described by Fabre and Delaloye (1983) and Moreno and
dagrande Complex marginal basin (Nivia et al., 1996; Concha (1993) and some volcanic input within Cretaceous
Nivia and Gómez, 2005). shales (Rubiano, 1989; Villamil, 1994).
Tectonic subsidence in the UMV (region 6, Fig. 6), Small mafic intrusions described by Fabre and Delaloye
where marine sedimentation started in Aptian times, is sig- (1983) coincide with areas of thin crust (crustal stretching
nificantly less than that of the EC and Middle MV (region factors >1.4) and places of maximum stretching of the sub-
2). In the easternmost LLA (region 10, Fig. 6), sedimenta- crustal lithosphere (Fig. 7c). As a consequence of the
tion started during the Late Cretaceous, and total tectonic depth-dependent lithosphere rheology assumed by the mod-
subsidence during the Cretaceous was small compared with el, the results suggest that more intense stretching affected the
the EC and MV. Subsidence probably occurred through subcrustal mantle lithosphere (Fig. 7d). Differences between
flexural thermal subsidence (Watts et al., 1982) and water crustal and subcrustal stretching factors suggest some decou-
loading due to an increase in paleowater depth. pling occurred between the crust and the subcrustal litho-
sphere or that increased thermal thinning affected the
7.2. Latest Jurassic–Early Cretaceous fast subsidence and mantle lithosphere. The latter interpretation implies a con-
lithosphere stretching events siderable thermal anomaly produced by mantle lithosphere
thinning, which seems supported by the presence of magmat-
7.2.1. Latest Jurassic–Hauterivian event (variable in ic mafic intrusions. During rifting, stress-induced litho-
different stratigraphic columns, 144–127 Ma) sphere thinning causes adiabatic decompression of the
This event occurred in the area of the EC and is best rep- lower lithosphere and asthenosphere, their partial melting,
resented in its eastern flank and west of the Bucaramanga and the diapiric rise of melts into the zone of thinned litho-
fault. Subsidence curves (Fig. 10), thickness maps (Figs. sphere (Wilson, 1993). Although the 1D model cannot pre-
7a, b, and 8a), paleomagnetic data (Bayona et al., 2005), dict regional isostatic effects, the Lower Cretaceous
and the distribution of crustal d stretching factors unconformity on the rift margins (e.g., LLA) and locally
(Fig. 7c) evidence a wide (>180 km), asymmetrical, trans- on horst blocks (e.g., Santander-Floresta paleo-Massif)
tensional half-graben basin divided by the Santander-Flor- probably was produced by thermal uplift of rift shoulders,
esta high. Maximum tectonic subsidence and crustal as suggested by the subcrustal stretching values. In addition,
stretching d up to 1.66 was associated with the pre-Guaicá- Jurassic magmatic activity in the Santander Massif contrib-
ramo normal master fault system, the eastern boundary of utes to relative thermal uplift and reduces tectonic subsi-
the graben (Figs. 7a and c). Strong changes in the thickness dence. According to Ziegler (1994), unconformities on rift
of the Girón Formation from 0 to more than 4 km across shoulders and intrabasinal fault blocks can be attributed to
ENE-WSW- and NE-SW-striking normal faults and local a footwall uplift in response to extensional unloading of
vertical axis rotations of fault-bounded blocks during this the lithosphere. In general, the location of subcrustal and
time, as documented by paleomagnetism, suggest an impor- crustal stretched zones coincides, as a consequence of the
tant control of normal faulting on deposition and basin 1D model assumption of local isostasy. However, where
configuration during the deposition of continental beds of there is some offset, it indicates asymmetry in the basin, as
the Girón Formation (Bayona et al., 2005; Fig. 10, supported by the general geometry of the basin. On the basis
Tablazo-Lebrija). A second-order half-graben was located of subsidence analyses of Cretaceous stratigraphic columns
at the current location of the western flank of the EC with of the EC, Fabre (1987) and Hébrard (1985), using the
crustal stretching values up to 1.45 (Fig. 7c). This minor instantaneous stretching model of McKenzie (1978), calcu-
half-graben probably was associated with a paleonormal late uniform b = d stretching factors up to 2 for the whole
fault system, approximately following La Salina-Bituı́ma lithosphere. However, they lump the Cretaceous stretching
fault system that was its western border. To the south, there events into a single ‘‘instantaneous’’ stretching event with
was only one depocenter, limited to the south by a NW-SE an infinite extension rate. The higher stretching values they
vertical transfer fault. Early Cretaceous turbidites at both obtain is a logical consequence of combining several stretch-
flanks (Murca Formation, Cáqueza Group) of the exten- ing events with finite extension rates.
sional basin can be taken as evidence of tectonic instability According to the Cretaceous passive margin interpreta-
associated with normal faulting movement. Branquet tion (Pindell and Erikson, 1993), active opening of the
(1999) presents outcrop and seismic evidence of Cretaceous proto-Caribbean occurred north of Colombia and west of
402 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Fig. 13. Seismic sections in the Medina foothills area, along the eastern border of the Cundinamarca subbasin. Note normal fault evidence during the
Cretaceous (K) in the Guaicáramo paleofault system along the eastern border and contractional inversion of Cretaceous extensional faults during the
Paleogene evidenced by lateral changes of thickness of the Paleogene Carbonera and Mirador formations. Note thickness changes in the Cretaceous
sedimentary fill (Linares, 1996). Location of seismic line is shown in Fig. 2.

the paleo-Central Cordillera. Moreno and Pardo (2003) cor- and Pardo (2003), subduction of the western Quebrada-
relate this latest Jurassic–Early Cretaceous extensional grande proto-Caribbean crust beneath the Farallon plate
phase, which originated the proto-Caribbean basin accord- originated a magmatic arc. To the east, a passive margin dis-
ing to Pindell and Erikson (1993) reconstruction, with the tant from the magmatic arc prevailed in the eastern side of
Quebradagrande Complex marginal basin of Nivia et al. the Quebradagrande Basin and east of the paleo-Central
(1996) and Nivia and Gómez (2005). According to Moreno Cordillera (Moreno and Pardo, 2003, their Fig. 7). If such
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 403

an interpretation is valid for the Berriasian–Hauterivian, the nogenic clays or bentonites in the Cenomanian–Turo-
event may be the result of stretching in the study area, which nian stratigraphic interval (Villamil and Arango, 1998)
produced a failed-rifted arm related to a major opening of and Salada Member of the La Luna Formation (Patter-
the proto-Caribbean oceanic basin. However, the presence son, 1970, in Rubiano, 1989), as well as subaqueous vol-
of some Early Cretaceous plutonic rocks in the Central Cor- canic tuffs in La Frontera and La Luna formations in the
dillera east of the Quebradagrande Complex (i.e., lower Cre- MV (Restrepo-Pace, personal communication).
taceous K/Ar ages of the Samaná and Mariquita stocks, 8. Jurassic (185 Ma) and Cretaceous (P77 Ma) zircon fis-
Vesga and Barrero, 1978) is difficult to explain with a passive sion-track ages from the Central Cordillera (Toro
margin hypothesis. One possibility is that these ages are not et al., 1999; Gómez et al., 1999), evidence of uplift. In
reliable. An alternative plate tectonic reconstruction attri- the Ecuadorian Andes (Rivadeneira, 1996) and Central
butes Early Cretaceous extension of the paleo-EC to a back- Cordillera (Rodrı́guez and Rojas, 1985), uplift or defor-
arc basin contemporaneous with reduced magmatic activity mation have been suggested during the Late Cretaceous.
in the Central Cordillera. The proto-Caribbean eastern part
of the Quebradagrande Complex basin crust may have sub- 7.2.2. Aptian–early Albian event (121–102 Ma)
ducted beneath the Central Cordillera to develop a magmat- This fast subsidence event mainly occurred at the south-
ic arc. Although these alternative interpretations are ern part of the western flank of the EC and the UMV, indi-
controversial, the following evidence supports an EC basin cating asymmetry in the basin. In other localities, thermal
located behind a partially emerged, less subsiding paleo- subsidence is due to lithospheric cooling after prior litho-
Central Cordillera (magmatic arc?): spheric stretching events. During the Barremian–Aptian,
the basin extended to the south in the UMV (Figs. 8a
1. Early Cretaceous igneous intrusions in the Central Cor- and 10). Turbiditic deposits of Aptian age (Socotá Mem-
dillera east and west of the Quebradagrande Complex ber, Polanı́a and Rodrı́guez, 1978) indicate tectonic insta-
(e.g., San Diego, Cambumbia, and Mariquita stocks, bility associated with the rapid subsidence event. The
Restrepo et al., 1991) define one or two (?) magmatic isopach map (Fig. 8a) suggests a master normal fault sys-
arcs (produced by subduction of the W and E borders tem, located approximately at the present-day Bituima
of the Queberadagrande basin crust?). However, such Magdalena fault system, was active. In the UMV, a normal
magmatic arcs are not well defined. paleo-Chusma fault system probably also was active.
2. The presence in the western part of the Cundinamarca Crustal stretching factors up to 1.4 are associated with
subbasin of Lower Cretaceous sandstones with abun- the southern segment of the Bituima fault system and those
dant volcanic lithic fragments and feldspar derived from up to 1.2 with the UMV (Fig. 8b). Because of the depth-de-
a western detrital source area, as indicated by paleocur- pendent rheology assumed by the model, the results suggest
rent data (Murca Formation and Útica sandstone; Sar- that stretching affected the subcrustal mantle lithosphere
miento, 1989; Moreno, 1990, 1991). more strongly. Subcrustal stretching values reach 3.24 at
3. Volcanic lithic fragments in the mid-Cretaceous Cabal- the southwestern flank of the EC and 1.6 at the UMV
los Formation of the UMV (Guerrero et al., 2000). (Fig. 8c). Differences between crustal and subcrustal
4. Progressive westerly onlap terminations of Cretaceous stretching values suggest some decoupling between crust
carbonates on the basement, observed in seismic lines and subcrustal lithosphere or that an increased thermal
in the western border of the Cesar Valley, northern thinning affected the mantle lithosphere. These results
Colombia (in Mesozoic times, part of the EC basinal imply a thermal anomaly that probably is responsible for
area, Fig. 1; Audemard, 1991). rift shoulder uplift, as evidenced from fission-track data
5. Stratigraphical and petrographical evidence suggesting by Van der Wiel (1991) in the UMV and Garzón Massif
that during the Berriasian (?)–Valanginian, clastic sedi- (Fig. 1). Van der Wiel (1991) interprets these ages as related
ments near San Felix in the western flank of the Central to an orogenic event that affected the northwestern corner
Cordillera (between Romeral and Palestina faults) came of South America between 100 and 80 Ma. However, this
from erosion of uplifted areas with metamorphic rocks interpretation contradicts strong stratigraphic evidence of
and small tectonic blocks with plutonic rocks (Rodrı́- a subsiding basin at these localities. The zircon ages reflect
guez and Rojas, 1985). local uplift of faulted blocks located at rift margins, as
6. Cretaceous volaniclastic rocks in the Central Cordillera demonstrated by fission-track data from several rift basins
(Rodrı́guez and Rojas, 1985) consisting of mixtures of (Van der Beek, 1995). Van der Beek (1995) explains rift
pyroclastic and epiclastic fragments, probably derived margin uplift by mechanical support of rift flanks, resulting
from a magmatic arc. from an upward state of flexure. Evidence of magmatic
7. Relatively high concentration of volcanogenic clay min- activity in the basin is limited to some small mafic intru-
erals in Hauterivian–Barremian (0–30%), middle Albian sions (Fabre and Delaloye, 1983) and minor volcanic input
(0–21%), and Turonian (6–9%) shales of the Villeta (Rubiano, 1989; Villamil and Arango, 1998).
Group (Rubiano, 1989) and the Valanginian–Hauteri- Plate tectonic interpretations by Meschede and Frisch
vian Rosablanca Formation (Moreno, 1989, in Rubiano, (1998) assume the beginning of subduction of the Faral-
1989) in the Cundinamarca subbasin. Thin beds of volca- lon/Pacific plate under the Panama-Costa Rica arc west of
404 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Colombia during the Aptian–early Albian. The alternative In conclusion, Jurassic (and probably post-Triassic)
hypothesis of Pindell and Erikson (1993) assumes that dur- fast tectonic subsidence events are related to backarc to
ing the Aptian, the proto-Caribbean lithosphere began to intra-arc transtensional lithospheric stretching, with a pre-
subduct under the Amáime-Antilles arc, which was dominant dextral strike-slip component. Early–middle
approaching the western margin of northern South America. Cretaceous fast tectonic subsidence events seem related
to backarc lithosphere stretching behind a poorly devel-
7.2.3. Cenomanian subsidence event (98–93 Ma) oped, subduction-related magmatic arc. An increase in
This event occurred in the eastern flank of the EC. The distance from the backarc basin to the magmatic arc from
Late Cenomanian–Turonian global sea level maximum cor- the Jurassic to the Cretaceous has been observed else-
relates with it, suggesting that an increase in subsidence was where in basins that evolved from an initial intra-arc stage
driven by water load. However, the event affected only the to a later backarc stage (Smith and Landis, 1995). Late
eastern flank of the Cordillera, where the maximum Creta- Cretaceous subsidence was mainly thermal subsidence
ceous thickness appears. At other localities, thermal subsi- after the previous lithosphere stretching events, and local-
dence resulted from lithospheric cooling after lithospheric ized discrete subsidence events probably result from litho-
stretching events. Although the maximum flooding surface sphere flexural bending due to horizontal compressive
for the Cretaceous sediments of the EC is the Cenoma- stresses related to accretion of the oceanic terranes in wes-
nian–Turonian boundary (Villamil and Arango, 1998), it is tern Colombia.
Campanian in the eastern LLA (Fajardo et al., 1993; Cooper
et al., 1995; Fig. 5). If subduction of the Caribbean’s thick 8. Discussion
and buoyant lithosphere (Duncan and Hargraves, 1984;
Nivia, 1987; Hill, 1993; Meschede and Frisch, 1998) under 8.1. Relationships between Mesozoic rifting and magmatism
South America was inhibited, it initiated uplift of the Central
Cordillera, as evidenced by Late Cretaceous fission-track In the study area, abundant Upper Triassic–Lower
ages from the Central Cordillera (Gómez et al., 1999; Toro Jurassic volcanic rocks are associated with moderate
et al., 1999) and exerted horizontal stresses on the northwest- stretching factors (b = d up to 1.12). In contrast, the
ern margin of South America. Horizontal stresses can induce Cretaceous sedimentary record is almost devoid of volca-
local flexural lithosphere bending, which is maximal where nic rocks (only volcanic tuffs and minor mafic intrusions)
the lithosphere is weakest (Cloetingh, 1988; Cloetingh and and associated with higher stretching factors (b up to 3,
Kooi, 1992). This process probably enhanced the relative d up to 1.66). Thermal processes were more important
sea level rise, creating a maximum Cenomanian–Turonian than mechanical stretching during Late Triassic–Early
marine flooding surface in the depocenter of the EC, charac- Jurassic rifting than during Cretaceous extension. During
terized by weak lithosphere due to earlier stretching. In con- the Late Triassic–Early Jurassic, abundant volcanic rocks
trast, horizontal stress produced a submarine shallow water suggest a positive thermal anomaly in the lithosphere but
depth bulge in the LLA, which partially compensated for the moderate lithosphere stretching. Triassic–Jurassic uncon-
maximum eustatic signal. formities could have been produced by thermal uplift
(‘‘active rifting’’?). Thermal doming results from progres-
7.2.4. Maastrichtian–Paleocene event (variable in different sive thinning of the higher density mantle lithosphere
columns, 68–54.8 Ma) and its replacement by low-density asthenosphere (Bott,
This fast subsidence event affected the axial part of the 1992). In contrast, during the Cretaceous, the less abun-
EC, its eastern flank, and, locally, the westernmost part dant volcanic rocks, absence of tectonically controlled
of the LLA. This event correlates in time with deformation unconformities, and large amount of tectonic subsidence
and uplift in the Central Cordillera (Jaramillo, 1978, 1981; suggest an absence of thermal doming. Small mafic intru-
Cooper et al., 1995). sions coincident with places of maximum crustal and
All plate tectonic interpretations agree that during the mantle subcrustal stretching also suggest modest magma-
latest Cretaceous and probably Paleocene, accretion of tism as a consequence of the extension of the lithosphere
the Western Cordillera oceanic terranes along the Cau- (‘‘passive rifting’’). The abundant Late Triassic–Early
ca-Patia fault occurred, producing deformation and uplift Jurassic volcanic rocks vary in composition from felsic
of the Central Cordillera. Increased subsidence in the to mafic. Chemical analyses of La Quinta Formation vol-
axis of the Cundinamarca subbasin (Sabana de Bogotá) canic rocks indicate a calc-alkaline composition in the
could result from increased horizontal compressional AFM diagram and alkaline composition in the alkali-sil-
stress (Cloetingh, 1988; Cloetingh and Kooi, 1992) asso- ica diagram (Toussaint, 1995b). Chemical analyses of the
ciated with the collision of the oceanic terranes of wes- Saldaña Formation indicate a calc-alkaline composition,
tern Colombia and deformation and uplift of the probably generated in a backarc environment (Bayona
Central Cordillera. Development of normal faults in et al., 1994). The predominance of calc-alkaline composi-
the LLA area, as suggested by Kluth et al. (1997), could tions seems to suggest a convergent-related backarc
be the result of local tensional stresses in the developed extension rather than intracontinental rifting (Toussaint,
flexural bulge. 1995b).
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 405

Fig. 14. Event correlation between lithosphere stretching in the area of the EC and magmatic activity in the Central Cordillera. (1) Triassic, (2) Late
Triassic–Early Jurassic, (3) Middle Jurassic, (4) Early Cretaceous Berriasian–Hauterivian, and (5) Aptian events. This correlation should be considered
preliminary because original data are heterogeneous; 94% are K-Ar (biotite, hornblende, muscovite, or whole-rock), and 6% are Rb-Sr (hornblende/biotite
or whole-rock). (a) Left panel: principal structural/plutonic zones of western Colombia. Right panel: age distribution of Mesozoic and Cenozoic plutonic
activity in western Colombia (modified after Aspden et al., 1987). (b) Cumulative histogram of radiometric ages of plutonic bodies in Colombia (modified
after Guillande, 1988). Periods of intense magmatic activity are characterized by rapid increase in the cumulative number of radiometric age
determinations for a time interval (low slope).

8.2. Correlation between fast subsidence events and backarc basins develop when the velocity rollback, due to
subduction-related magmatic arcs fast subduction, exceeds the oceanward convergence
velocity of the overriding plate (Royden, 1993a,b). If
The inferred Mesozoic stretching events seem to corre- magmatic arc activity decreases with the oceanward con-
late in time with reduced magmatic activity in the Central vergence velocity of the overriding plate, during times of
Cordillera (Fig. 14, modified from Aspden et al., 1987; reduced magmatic arc activity, a constant rollback veloc-
Guillande, 1988). However, this preliminarily correlation ity exceeds that velocity, increasing extension and subsi-
should be tested with more and better data. If the calc- dence in the backarc region. According to Aspden et al.
alkaline (Alvarez, 1983) plutonic belts of the Central (1987), the Triassic magmatic belt was controlled along
Cordillera and Santander Massif developed as subduc- strike-slip faults, as also supported by Restrepo-Pace
tion-related magmatic arcs during Mesozoic times, as sug- (1995). The Jurassic magmatic arc may have been con-
gested by Aspden et al. (1987), the extensional or trolled by strike-slip faults, as suggested by paleomagnetic
transtensional basins behind them or close to them may data (Bayona et al., 2005) and its elongated shape in
be interpreted as backarc or intra-arc basins. Extensional map view, parallel to major strike-slip faults of major
406 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

calc-alkaline plutonic bodies (i.e., Ibagué and Segovia land et al. (1994). These Ecuadorian terranes were affected
batholiths). Along these faults, part of the magmatic belt by intense ductile deformation and transcurrent faulting
might have moved northward, as suggested by paleomag- (Litherland et al., 1994). The early Mesozoic rift systems
netic data (Bayona et al., 2005). Jurassic calc-alkaline plu- of the Payandé-San Lucas terranes in Colombia and the
tonism and volcanism along the Central Cordillera and Peruvian/Bolivian EC represent transtensional segments
UMV is interpreted by Aspden et al. (1987) and Bayona separated by a transpresional ‘‘node’’ in Ecuador, in line
et al. (1994) as a subduction-related magmatic arc, but with the map view geometry of a major dextral strike-slip
Cretaceous plutonism developed sporadically only in the fault system. In Colombia, the calc-alkaline magmas likely
northern part of the Central Cordillera (Restrepo et al., intruded along active strike-slip faults and shear zones,
1991), whereas it is very extensive in Peru (Cobbing, developing a magmatic arc that may have resulted from a
1982, in Aspden et al., 1987). Aspden et al. (1987) suggest highly oblique subduction zone.
oblique convergence and an offset in the subduction zone Unlike the calc-alkaline magmatism of the Colombian
along a major NE-SW transform fault to account for the Triassic–Jurassic magmatic arc segments, the Late Perm-
notable absence of Cretaceous plutonism in southern ian–Middle Jurassic rift system of the EC of Peru and
Colombia and Ecuador. Bolivia is characterized by intense magmatism, predomi-
nantly alkaline to tholeiitic, probably related to intraplate
8.3. Geometry of transtensional basins magmatism and lithospheric thinning (Sempere et al.,
2002). In southern Peru and Bolivia, the Late Permian–
8.3.1. Triassic and Jurassic Middle Jurassic rift system is located behind the Jurassic
On the basis of paleomagnetic data, Bayona et al. (2005) magmatic arc (Sempere et al., 2002, Fig. 3). In central Peru,
suggest that the Payandé terrane (UMV) was located south plutons probably were emplaced in the rift roots (Sempere
of equator (close to 10 S) before the Middle Jurassic; Sem- et al., 2002). The Colombian Payandé-San Lucas terranes
pere et al. (2002) recognize Late Permian–Middle Jurassic may have originated as fragments of thinned lithosphere
transtensional rifting in the EC of Peru and Bolivia (8– separated from the Peru–Bolivia rift system by intraplate
22 S). The early Mesozoic stratigraphy of the Peruvian– rifting, which later moved north along strike-slip faults.
Bolivian rift basin is very similar to that of the UMV The Late Permian–Middle Jurassic rift system of the EC
(Payandé terrane, sensu Etayo-Serna et al., 1983). In Peru of Peru and Bolivia is interpreted by Sempere et al.
and Bolivia, Late Permian–Triassic, red to purple, coarse- (2002) as a transcurrent rift system in which transtensional
grained, continental, synrift deposits (Mitu Group) resem- segments were separated by transpressional nodes. This
ble those of the Luisa Formation of the UMV; Late transcurrent setting also could explain Triassic plutons in
Triassic–Early Jurassic (Norian-Liassic), marine carbon- Peru and Bolivia with deformation contemporaneous to
ate-dominated postrift deposits (Pucara Group) resemble their emplacement (Sempere et al., 2002), as well as then-
those of the Payandé Formation of the UMV; and Jurassic, foliated Triassic plutons described in Colombia (Aspden
brown-red mudstones and coarse-grained alluvial deposits et al., 1987). However, new paleomagnetic, geochronolog-
(Sarayaquillo Formation) resemble the sedimentary com- ical, isotopic, petrological, structural, and stratigraphic
ponent of the Saldaña Formation of the UMV. The simi- research in necessary to test this hypothesis.
larity of the early Mesozoic stratigraphy in these areas
initially was recognized by Geyer (1982). Synrift deposits 8.3.2. Cretaceous
of Peru and Bolivia commonly are interbedded with locally During the Cretaceous, the Colombian transtensional
dominant volcanic and volcaniclastic rocks and/or intrud- basins increased in width. During the Mesozoic, the
ed by subvolcanic to plutonic rocks (Sempere et al., 2002). strike-slip component gradually decreased at the expense
A similar association with volcanic, volcaniclastic, and plu- of increases in the extensional component, as preliminarily
tonic rocks is observed in the UMV. Paleomagnetic data suggested by paleomagnetic data (Bayona et al., 2005) and
(Bayona et al., 2005) and the similarity of the early Meso- lithosphere stretching values. As in other areas (e.g., East
zoic stratigraphy and basin tectonic setting of the Payandé African rift), probably during the initial Cretaceous exten-
terrane and Peru–Bolivia rift probably indicate that both sional stages, reactivation of crustal discontinuities led to
areas were adjacent before the Middle Jurassic. In Peru the subsidence of isolated grabens linked by shear zones.
and Bolivia, rifting events occurred during the Late Perm- The increase in width probably was the result of increasing
ian–Middle Jurassic interval (Sempere et al., 2002) and sep- strain, with grabens propagating toward one another, coa-
arated during the Late Triassic–Early Jurassic, thermal sag, lescing, and evolving into a generally continuous rift sys-
postrift event during deposition of the Pucara Group (Sem- tem (Ziegler, 1994). Whether the Colombian Mesozoic
pere et al., 2002). Similar punctuated rift and lithosphere extensional basins were pure shear or simple shear exten-
stretching events occurred in the UMV (Payandé terrane). sional basins is difficult to demonstrate. Probably both
Geographically, however, between these Colombian and mechanisms operated; these extensional basin models
Peruvian/Bolivian transtensional basins in Ecuador during should be viewed as end-member cases. If the orientation
Jurassic, the accretion of terranes preserved as highly of preexisting crustal discontinuities is such that they could
deformed metamorphic rocks has been proposed by Lither- not be reactivated by the stress system governing the
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 407

evolution of the transtensional basin, new faults would (Fig. 7c, Sarmiento, 1989) probably represent Cretaceous
develop, and pure shear deformation will likely prevail transfer faults. If a subduction-related magmatic arc existed
(Ziegler, 1994). This mechanism may be applicable to the at the current location of the Central Cordillera during the
Triassic–Jurassic extensional basin system in Colombia. Cretaceous, as has been proposed (Aspden et al., 1987; Tous-
However, if the upper crust is weakened by the presence saint and Restrepo, 1989, 1994; Toussaint, 1995b), and the
of preexisting crustal discontinuities and is oriented to favor orientation of that magmatic arc and the extensional basins
reactivation under the prevailing tensional stress field, they has been preserved, their oblique orientation and en échelon
will present zones of preferential strain concentration, pattern would suggest oblique slip extension with a right-lat-
which can result in simple shear deformation (Ziegler, eral strike-slip component. However, the available data can-
1994). This mechanism may explain the development of not rule out the hypothesis that some rift arms form acute
the paleo-EC rift system during the Early Cretaceous: The angles to the dominant NNE-SSW trend in a pattern similar
eastern side of the extensional basin developed during the to aborted aulacogen rifts.
Berriasian–Hauterivian by reactivation of an older Paleozo- If the inverse or thrust faults that now define the eastern
ic rift system along the Guaicáramo paleofault (Hossack and western borders of the EC originally were Cretaceous
et al., 1999), and the western side developed by reactivation normal faults with a strike-slip component, inverted during
of earlier transtensional fault systems from the Triassic– the Cenozoic, their geometry in map view would provide
Jurassic. Another alternative is that the eastern and western some information about Mesozoic extensional faults. Lat-
margins of the graben developed along older strike-slip eral changes of Mesozoic thickness suggest this is the case,
fault systems, which bounded the accretion of tectonic terr- at least for the master faults, which probably defined the
anes west of the Guaicáramo paleofault system during the regional extensional basin geometry. Adopting this hypoth-
Late Triassic and Jurassic (Bayona et al., 2005). esis, we posit that the Guaicáramo, La Salina, Bitúima, Mag-
The rheological properties of the lithosphere control the dalena, and Boyacá faults represent original extensional or
depth at which tensional necking occurs and whether a rift transtensional faults. Analog model experiments of oblique
zone is flexed up- or downward (Ziegler, 1994). A deep lith- extension produce a similar map view fault pattern (e.g.,
osphere necking level causes upward flexure of the rift Tron and Brun, 1991). NW-SE transfer faults and possible
zone, whereas necking at shallow crustal levels causes normal faults with this orientation, as interpreted by Ecope-
downward flexure and the absence of shoulder uplifts trol (1994) in the middle MV, were not inverted during the
(Kooi et al., 1992; Ziegler, 1994). Similar deep levels of Cenozoic. Some normal faults were passively transported
necking in the eastern side of the Early Cretaceous exten- with short-cut basement blocks during the Cenozoic inver-
sional basin system may have generated shoulder uplift in sion (e.g., Esmeraldas fault; Cooper et al., 1995).
the LLA area during the Early Cretaceous. Coarse detrital
fragments in the Lower Cretaceous Brechas de Buenavista 9. Conclusions
(Pimpirev et al., 1992) and Calizas del Guavio (Conglome-
rado de Miralindo, Ulloa and Rodrı́guez, 1976) formations High-resolution backstripping analysis and forward
could be derived from this graben shoulder. In contrast, in modeling show that the Mesozoic Colombian Basin is
the western margin of the Early Cretaceous extensional marked by five lithosphere stretching pulses. Periods of
basin, sedimentation was more continuous from Jurassic extension seem to correlate in time with gaps of subduc-
to Early Cretaceous times, which implies the downward tion-related magmatic arc activity, as suggested by Aspden
flexure of the rift shoulders and thus a shallower level of et al. (1987), especially during the Jurassic, which supports
necking during Early Cretaceous times. a hypothesis of backarc extension. However, this prelimi-
On a lithospheric scale, the location of rift systems is con- narily correlation must be tested with more and better geo-
trolled by the location of weakness zones in the lithosphere, chronological and biostratigraphical data. If backarc
which in turn depends on its thermal state and crustal thick- extension continued during the Early Cretaceous through
ness. At crustal scales, the composition, thickness of its oblique plate convergence, it probably has a strong
mechanically strong upper layer, and availability of internal strike-slip component. Evidence of a backarc basin located
discontinuities (which can tensionally be reactivated) are behind a partially emerged, less subsiding paleo-Central
also important controls for rift location (Ziegler, 1994). Cordillera (magmatic arc?) during the Late Jurassic–Early
The overall present-day pattern of these transtensional Cretaceous includes the following: (1) Cretaceous plutonic
basins for the Cretaceous indicates several grabens oriented bodies in the Central Cordillera; (2) Lower Cretaceous
NNE-SSW in an en échelon pattern, in contrast with the sandstones in the western part of the Cundinamarca subba-
more N-S–oriented Central Cordillera (e.g., Mojica et al., sin with abundant volcanic lithic fragments and feldspar
1996). Some authors (Fabre, 1987; Sarmiento, 1989; Geotec, derived from a western detrital source area, indicated by
1992; Mojica et al., 1996) suggest that some NW-SE faults paleocurrent data (Murca Formation and Útica sand-
probably represented transfer faults. Features such as the stone); (3) progressive westerly onlap terminations of the
Nazareth NW-SE fault (Fig. 7c), which limits the Early Cre- Cretaceous carbonates on the basement, observed in seis-
taceous basin to the south (Fabre, 1987), or the NW-SE mic lines, in the western border of the Cesar Valley in
alignment connecting the two emerald districts of the EC northern Colombia; (4) petrographical evidence suggesting
408 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

that Berriasian (?)–Valanginian clastic sediments near San increase of the extensional component, according to
Felix in the western flank of the Central Cordillera came paleomagnetic data (Bayona et al., 2005) and lithosphere
from the erosion of nearly uplifted areas containing frag- stretching values. During the Berriasian–Hauterivian, the
ments of metamorphic rocks and small tectonic blocks with eastern side of the extensional basin may have developed
plutonic rocks; (5) Cretaceous volaniclastic rocks that by reactivation of an older Paleozoic rift system associat-
probably also derived from a magmatic arc; and (6) Late ed with the Guaicáramo fault system (Hossack et al.,
Cretaceous zircon fission-track ages in the Central Cordil- 1999). The western side probably developed through
lera. However, a passive margin hypothesis (Pindell and reactivation of an earlier normal fault system developed
Erikson, 1993) or aborted rift arms related to the Caribbe- during Triassic–Jurassic transtension. Alternatively, the
an opening cannot be completely ruled out because of the eastern and western margins of the graben may have
absence of a well-defined Cretaceous magmatic arc. developed along older strike-slip faults, the boundaries
Preliminarily, three stretching events are suggested dur- of the accretion of terranes west of the Guaicáramo fault
ing Triassic–Jurassic times. These events must be consid- during the Late Triassic and Jurassic (Bayona et al.,
ered preliminary because the poor continental Triassic 2005). During the first stages of Cretaceous lithosphere
and Jurassic fossiliferous sedimentary record offers rela- extension, lithosphere necking started at a deep level,
tively scarce, low-quality biostratigraphic data that make then evolved to shallower levels in latter extensional stag-
defining clear time boundaries between events difficult. es. The increasing width of the graben system likely was
The spatial distribution of the lithosphere stretching values the result of progressive tensional reactivation of preex-
suggests that small, narrow (<150 km), asymmetric graben isting upper crustal weakness zones. Lateral changes of
basins were located on opposite sides of the paleo-Magda- Mesozoic sediment thickness suggest that the reverse or
lena–La Salina fault system, which probably was active as thrust faults that now define the eastern and western
a master transtensional or strike-slip fault system. Paleo- borders of the EC originally were normal faults with a
magnetic data suggesting a significant (at least 10) north- strike-slip component that inverted during the Cenozoic
ward translation of terranes west of the Bucaramanga fault Andean orogeny. Thus, the Guaicáramo, La Salina,
during the Early Jurassic (Bayona et al., 2005) and the sim- Bitúima, Magdalena, and Boyacá were originally trans-
ilarity of early Mesozoic stratigraphy and the tectonic set- tensional faults. The oblique orientation of most relative
ting of the Payandé terrane with the Late Permian to the Mesozoic magmatic arc of the Central Cordillera
transtensional rift of the EC of Peru and Bolivia (Sempere may result from oblique slip extension during the Creta-
et al., 2002) suggest that the areas may have been adjacent ceous or was inherited from the pre-Mesozoic structural
in early Mesozoic times. Additional geochronological, pet- grain. However, not all Mesozoic transtensional faults
rological, stratigraphic, and structural research should ver- were inverted; some normal faults were passively trans-
ify this hypothesis. In particular, additional paleomagnetic ported with short-cut basement blocks during Cenozoic
investigations should determine the paleolatitudinal posi- inversion.
tion of the Central Cordillera and adjacent tectonic terr- Thermal processes were more dominant than mechani-
anes during the Triassic–Jurassic. Subsidence curves, cal stretching during Late Triassic–Early Jurassic phases
isopach maps, and paleomagnetic data (Bayona et al., than the Cretaceous extensional phase. During the Late
2005) suggest that during the Berriasian–Hauterivian, a Triassic–Early Jurassic, abundant volcaniclastic rocks sug-
(>180 km) wide, asymmetrical, transtensional half-rift gest a positive thermal anomaly in the lithosphere but mod-
basin existed, divided by the Santander Floresta high. A erate lithosphere stretching. Triassic–Jurassic age
single depocenter in the south was limited at its southern unconformities could have been produced by thermal uplift
reach by a vertical transfer fault. Small mafic intrusions (active rifting?). In contrast, during the Cretaceous, less
coincide with areas of thin crust (crustal stretching factors abundant volcanic rocks, the absence of tectonically con-
>1.4) and places of maximum stretching of the subcrustal trolled unconformities, and significant tectonic subsidence
lithosphere. During the Aptian–early Albian, the basin indicates the absence of thermal doming. Minor mafic
extended south in the UMV. Different crustal and sub- intrusions coincident with places of maximum crustal and
crustal stretching values suggest either some lowermost mantle subcrustal stretching suggest modest magmatism
crustal decoupling between crustal and subcrustal litho- took place as a result of the extension of the lithosphere
sphere or increased thermal thinning that affected the man- (passive rifting).
tle lithosphere. Late Cretaceous subsidence was mainly
driven by lithospheric cooling, water loading, and horizon- Acknowledgments
tal compressional stresses generated by collision of oceanic
terranes in western Colombia. This research was part of the doctoral thesis of L.F. Sar-
Triassic transtensional basins were narrow and miento, supported in Colombia by the Icetex-Ecopetrol
increased in width during Triassic and Jurassic times. Fund, the Petroleum Colombian Company ECOPETROL,
Cretaceous transtensional basins were wider than Trias- and the Instituto Colombiano del Petróleo ICP and in the
sic–Jurassic basins. During the Mesozoic, the strike-slip Netherlands by the Netherlands Research School of Sedi-
component gradually decreased at the expense of the mentary Geology and the Tectonics Group of the Vrije
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 409

Universiteit at Amsterdam. L.F. Sarmiento publishes with A.J., Taborda, A., 1995. Basin development and tectonic history of the
the permission of Ecopetrol-ICP. We acknowledge German Llanos Basin, Eastern Cordillera, and Middle Magdalena Valley,
Colombia. A.A.P.G. Bull. 79 (10), 1421–1443.
Bayona, who shared his paleomagnetic results that comple- Duncan, R.A., Hargraves, R.B., 1984. Plate tectonic evolution of the
mented a previous version of this article and suggested Caribbean region in the mantle reference frame. In: Bonini, W.E.,
improvements to the manuscript. This article also benefited Hargraves, R.B., Shagam, R. (Eds.), The Caribbean-South American
from reviews by François Roure, Reini Zoetemeijer, Harry Plate Boundary and Regional Tectonics. Geol. Soc. Am., Mem. 162,
Doust, and Pedro Restrepo-Pace. 81–93.
Ecopetrol, Esso and Exxon Exploration Company, 1994, Integrated
technical evaluation Santander sector Colombia, 1991–1994. Report,
References text and figures. Final Report, Houston, 38p., 49 figs.
Etayo-serna, F., Renzoni, G., Barrero, D., 1976. Contornos sucesivos del
Alvarez, J., 1983. Geologı́a de la Cordillera Central y el Occidente mar Cretácico en Colombia. In: Primer Congreso Colombiano de
Colombiano, y petroquı́mica de los intrusivos granitoides Mesoce- Geologı́a, Mem., Bogotá, pp. 217–252.
nozóicos. Ph.D. Thesis Univ. Chile. Boletin Geológico Ingeominas, Etayo-Serna, F. (Ed.), 1994. Estudios geológicos del Valle Superior del
Bogotá 26(2), 1–175. Magdalena. Univ. Nacional de Colombia, Ecopetrol, Bogotá, 200p.
Aspden, J.A., McCourt, Brook, M., 1987. Geometrical control of Etayo-Serna, F., Laverde-Montaño, F. (Eds.), 1985. Proyecto Crétacico,
subduction-related magmatism: the Mesozoic and Cenozoic plutonic contribuciones. Ingeominas, Publ. Geol. Esp. 16, Bogotá, 200p.
history of Western Colombia. J. Geol. Soc. London 144, 893–905. Etayo-Serna, F., Barrero, D., Lozano, H., Espinosa, A., González, H.,
Audemard, F., 1991. Tectonics of Western Venezuela. Ph.D. Thesis. Rice Orrego, A., Zambrano, F., Duque, H., Vargas, R., Núñez, A., Álvarez,
University, Houston, TX, 243p. J., Ropaı́n, C., Ballesteros, I., Cardozo, E., Forero, H., Galvis, N.,
Bayona, G.A., Garcı́a, D.F., Mora, G., 1994. La Formación Saldaña: Ramı́rez, C., Sarmiento, L., 1983. Mapa de terrenos geológicos de
producto de la actividad de estratovolcanes continentales en un Colombia. Ingeominas Publ. Geol. Esp. 14, Bogotá, 235p.
dominio de retroarco. In: Etayo Serna, F. (Ed.), Estudios geológicos Fabre, A., 1983a. La subsidencia de la Cuenca del Cocuy (Cordillera
del Valle Superior del Magdalena. Univ. Nacional de Colombia, Oriental de Colombia) durante el Cretáceo y el Terciario Inferior.
Ecopetrol, Bogotá, 21p., Chapter I. Primera parte: Estudio cuantitativo de la subsidencia. Geologı́a
Bayona, G., Rapalini, A.E., Constanzo-Alvarez, V., Montes, C., Veloza, Norandina, Bogotá 8, 49–61.
G., Ayala-Calvo, R.C., Gómez-Casallas, M., Silva, C., 2005. Mesozoic Fabre, A., 1983b. La subsidencia de la Cuenca del Cocuy (Cordillera
terrane translations and crustal block rotations in the Eastern Oriental de Colombia) durante el Cretáceo y el Terciario Inferior.
Cordillera and Magdalena Valley of Colombia, as inferred from Segunda parte: Esquema de evolución tectónica. Geologı́a Norandina,
paleomagnetism. In: 6th International Symposium of Andean Geody- Bogotá 8, 21–27.
namics, 4p., submitted. Fabre, A., 1985. Dinámica de la sedimentación Cretácica en la región de la
Bessis, F., 1986. Some remarks on the study of subsidence of sedimentary Sierra Nevada del Cocuy (Cordillera Oriental de Colombia). In:
basins; application to the Gulf of Lions margin (western Mediterra- Etayo-Serna, F., Laverde-Montaño, F. (Eds.), Proyecto Crétacico,
nean). Marine Petroleum Geol. 3 (1), 37–63. contribuciones. Chapter XIX, Ingeominas Publ. Esp. 16, Bogotá, 20p.
Bond, G., Kominz, M.A., 1984. Construction of tectonic subsidence Fabre, A., 1987. Tectonique et géneration d’hydrocarbures: Un modèle de
curves for the early Paleozoic miogeocline, southern Canadian Rocky l’évolution de la Cordillère Orientale de Colombie et du Bassin des
Mountains: implications for subsidence mechanisms, age of break-up, Llanos pendant le Crétacé et le Tertiaire. Arch. Sci. Genève 40 (Fasc.
and crustal thinning. Geol. Soc. Am. Bull. 95, 155–173. 2), 145–190.
Bott, M.H.P., 1992. Passive margins and their subsidence. J. Geol. Soc. Fabre, A., Delaloye, M., 1983. Intrusiones básicas Cretácicas de la
London 149, 805–812. Cordillera Oriental. Geologı́a. Norandina, Bogotá 6, 19–28.
Branquet, Y., 1999. Etude structurale et métallogénique des gisements Fajardo, A., Rubiano, J.L., Reyes, A., 1993. Estratigrafı́a de secuencias de
d 0 émeraude de Colombie: Contribution à l 0 histoire tectono-sédimen- las rocas del Cretáceo Tardı́o al Eoceno Tardı́o en el sector central de
taire de la Cordillère Orientale de Colombie. Thèse de Doctorat de la cuenca de los Llanos Orientales. Departamento del Casanare,
l’Institut National Polytechnique de Lorraine, Specialité Geosciences, Report ECP-ICP-001-93, Piedecuesta, Santander, 69p.
Matières premières et environement, Institut National Polytechnique Föllmi, K.B., Garrison, R.E., Ramirez, P.C., Zabrano-Ortı́z, Kennedy,
de Lorraine, Centre de recherches pétrographiques et géochimiques W.J., Lehner, B.L., 1992. Cyclic phosphate-rich successions in the
(CRPG/CNRS), 265p. Upper Cretaceous of Colombia. Palaeogeography, Palaeoclimatology,
Bürgl, H., 1961. Historia geológica de Colombia. Rev. Acad. Col. Cienc. Palaeoecology 93, 151–182.
Exac. Fis. Nat. Bogotá 9 (43), 137–191. Geotec, 1992. Facies distribution and tectonic setting through the
Bürgl, H., 1967. The orogenesis of the Andean system of Colombia. Phanerozoic of Colombia. A regional synthesis combining outcrop
Tectonophysics 4 (4-6), 429–443. and subsurface data presented in 17 consecutive rock-time slices.
Cediel, F., 1968. El Grupo Girón, una molasa Mesozóica de la Cordillera Bogotá, 100p.
Oriental. Serv. Geol. Nal. Bol. Geol. Bogotá 16 (1–3), 5–96. Geyer, O., 1982. Comparaciones estratigráficas y faciales en el Triásico
Cediel, F., Shaw, R.P., Cáceres, C., 2003. Tectonic assembly of the Norandino. Geologı́a. Norandina, Bogotá 5, 27–31.
Northern Andean Block. In: Bartolini, C., Buffer, R.T., Blickwede, J. Gómez, E., Pedraza, P., 1994. El Maastrichtiano de la región Honda
(Eds.), The Circum-gulf of Mexico and the Caribbean: Hydrocarbon Guaduas, lı́mite N del Valle Superior del Magdalena: Registro
Habitats, Basin Formation, and Plate Tectonics. AAPG memoir 79, sedimentario de un delta dominado por rı́os trenzados. In: Etayo
pp. 815–848. Serna, F. (Ed.), Estudios geológicos del Valle Superior del Magdalena.
Cloetingh, S., 1988. Intra-plate stress: a new element in basin analysis. In: Chapter III, Univ. Nacional de Colombia, Ecopetrol, Bogotá, 20p.
Kleinspehn, K.L., Paola, C. (Eds.), Frontiers in sedimentary geology – Gómez, E., Jordan, T., Hegarty, K., Kelley, S., 1999. Diachronous
New perspectives in basin analysis. Springer-Verlag, New York, NY, deformation of the Central and Eastern Andean cordilleras of Colombia
pp. 205–230. and syntectonic sedimentation in the Middle Magdalena Valley Basin.
Cloetingh, S., Kooi, H., 1992. Intraplate stress and dynamical aspects of In: Fourth ISAG, Gottingen, Germany, 4–6 October, pp. 287–290.
rift basins. In: Ziegler, P.A. (Ed.), Geodynamics of Rifting, Vol. II. Gradstein, F.M., Ogg, J.G., 1996. A Phanerozoic time scale. Episodes 9
Thematic Discussions. Tectonophysics 215, 167–185. (1-2), 3–5.
Cooper, M.A., Addison, F.T., Alvarez, R., Coral, M., Graham, R.H., Guerrero, J., Sarmiento, G., Navarrete, R.E., 2000. The stratigraphy of
Hayward, A.B., Howe, S., Martinez, J., Naar, J., Peñas, R., Pulham, the W side of the Cretaceous Colombian Basin in the Upper
410 L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411

Magdalena Valey. Revaluation of the select areas andtype localities Moreno, J.M., Concha, A.E., 1993. Nuevas manifestaciones ı́gneas
including Aipe, Guaduas and Piedras. Geologı́a Colombiana. Bogotá básicas en el flanco occidental de la Cordillera Oriental, Colombia.
25, 45–110. Geologı́a Colombiana 18, 143–150.
Guillande, M.R., 1988. Evolution Méso-Cénozoı̈que d’une vallée inter- Moreno, M., Pardo, A., 2003, Stratigraphical and sedimentological
cordilléraine andine: La Haute Vallée du Rio Magdalena (Colombie). constraints on Western Colombia: implications on the evolution of the
Ph.D. Thesis Univer. Paris 6, Paris, 352p. Caribbean Plate. In: Bartolini, C., Buffer, R.T., Blickwede, J. (Eds.), The
Hébrard, F., 1985. Les foothills de la Cordillère Orientale de Colombie Circum-gulf of Mexico and the Caribbean: Hydrocarbon Habitats,
entre les rios Casanare et Cusiana. Evolution géodynamique depuis Basin Formation, and Plate Tectonics. AAPG memoir 79, 891–924.
d’Eo Crétacé. Ph.D. Thesis. Univ. Pierre et Marie Curie, Paris, 162p. Nivia, A., 1987, The geochemistry and origin of the Amáime and volcanic
Hill, R.I., 1993. Mantle plumes and continental tectonics. Lithos 30, 193– sequences, SW Colombia. M. Phil. Thesis, Univ. of Leicester, UK,
206. 164p.
Hossack, J., Martı́nez, J., Estrada, C., Herbert, R., 1999. Structural Nivia, A., Gómez, J., 2005, Consideraciones acerca del modelo geológico
evolution of the Llanos fold and thrust belt, Colombia. In: McClay, K. evolutivo del Occidente Colombiano (Colombia). Resúmenes X
(Ed.), Thrust Tectonics 99 Meeting. Royal Halloway Univ. of London, Congreso Colombiano de Geologı́a. Bogotá, 2p.
April 26–29, June 1999, Program, London, 110p. Nivia, A., Mariner, G., Kerr, A.C., 1996. El Complejo Quebradagrande
Ingersoll, R.V., Busby, C.J., 1995. Tectonics of sedimentary basins. In: una possible cuenca marginal intracratónica del Cretáceo Inferior en la
Busby, C.J., Ingersoll, R.V. (Eds.), Tectonics of Sedimentary Basins. Cordillera Central de los Andes Colombianos. In: Paper presented to
Blackwell Science, pp. 1–5. the VII Congreso Colombiano de Geologı́a, unpublished.
Jaramillo, J.M., 1978. Determinación de las edades de algunas rocas de la Pessagno, E.A., Martin, C., 2003. Tectonostratigraphic evidence for the
Cordillera Central de Colombia por el método de huellas de fisión. origin of the Gulf of Mexico. In: Bartolini, C., Buffer, R.T., Blickwede,
2do. In: Congreso Colomb. Geol. Resúmenes, Bogotá, pp. 19–20. J. (Eds.), The Circum-gulf of Mexico and the Caribbean: Hydrocarbon
Jaramillo, J.M., 1981. Determinación de las edades de algunas rocas de la Habitats, Basin Formation, and Plate Tectonics. AAPG memoir 79,
Cordillera Central de Colombia por el método de huellas de fisión. Bol. 46–74.
Cienc. de la Tierra, Medellı́n 5–6, 145–146. Pimpirev, C.T., Patarroyo, P., Sarmiento, G., 1992. Stratigraphy and
Julivert, M., 1968. In: Colombie. Lexique stratigraphique International. facies analysis of the Caqueza Group, a sequence of Lower Cretaceous
Precambrian, Paleozoique et Mesozoique. Paris, vol. V, Fasc. 4a, turbidites in the Cordillera Oriental of the Colombian Andes. Geol.
premiere partie, 651p. Colombiana, Bogotá 17, 297–308.
Kammer, A., 1993. Steeply dipping basement faults and associated Pindell, J., Dewey, J., 1982. Permo-Triassic reconstruction of western
structures of the Santander Massif, Eastern Cordillera, Colombian Pangaea and the evolution of the Gulf of Mexico-Caribbean region.
Andes. Geol. Colombiana, Bogotá 18, 47–64. Tectonics 1, 179–211.
Kluth, Ch., F., Laad, R., De Armas, M., Gómez, L., Tilander, N., 1997. Pindell, J., Erikson, J., 1993. The Mesozoic margin of northern South
Different structural styles and histories of the Colombian foreland: America. In: Salfity, J. (Ed.), Cretaceous tectonics of the Andes.
Castilla and Chichimene oil fields areas, east-central Colombia. In: VI Vieweg Germany, pp. 1–60.
Simpolsio Bolivariano Exploracion en las Cuencas Subandinas, Pindell, J.L., Tabbutt, K.D., 1995. Mesozoic-Cenozoic Andean paleoge-
Cartagena de Indias, Mem. Tomo II, pp. 185–197. ography and regional controls on hydrocarbon systems. In: Tankard,
Kooi, H., Burrus, J., Cloetingh, S., 1992. Lithospheric necking and A.J., Suarez, R., Welsink, H.J. (Eds.), Petreoleum Basins of South
regional isostasy at extensional basins: part 1, Subsidence and gravity America. A.A.P.G. Mem. 62, 101–128.
modeling with an application to the Gulf of Lions margin (SE France). Polanı́a, J.H., Rodrı́guez, O.G., 1978. Posibles turbiditas del Cretáceo
J.G.R. 97, 17553–17571. Inferior (Miembro Socotá) en el área de Anapoima (Cundinamarca);
Linares, R., 1996. Structural styles and kinematics of the Medina area, una investigación sedimentológica basada en registros gráficos. Geol.
Eastern Cordillera, Colombia. M.Sc. Thesis, Univ. of Colorado, 104p. Colombiana, Bogotá 10, 87–91.
Litherland, M., Aspden, J.A., Jamielita, R.A., 1994. The metamorphic belts Restrepo, J.J., Toussaint, J.F., Gonzalez, H., Gordani, U., Kawashita, K.,
of Ecuador. British Geological Survey, Overseas Memoir. 11, 1–146. Linares, E., Parica, C., 1991. Precisiones geocronológicas sobre el
Maze, W.B., 1984. Jurassic la Quinta Formation in the Sierra de Perijá, occidente Colombiano. Simposio sobre el magmatismo Andino y su
northwestern Venezuela: geology and tectonic environment of red beds marco tectónico. Programa Internacional de Correlación Geológica
and volcanic rocks. In: Bonini, W.E., Hargraves, R.B., Shagam, R. UNESCO, Union Internacional de Ciencias Geológicas. Manizales,
(Eds.), The Caribbean-South American Plate Boundary and Regional Colombia, Julio 3–5 de 1991. Manizales, pp. 1–21.
Tectonics. Geol. Soc. Amer. Mem. 162, 263–282. Restrepo-Pace, P.A., 1995. Late Precambrian to Early Mesozoic tectonic
McCourt, W.J., Feininger, T., Brook, M., 1984. New geological and evolution of the Colombian Andes, based on new geochronological,
geochronological data from the Colombian Andes: continental growth geochemical and isotopic data. Ph.D. Thesis. Univ. Arizona, 195p.
by multiple accretion. J. Geol. Soc. London 141, 831–845. Rivadeneira, M.V., 1996. Cretaceous-Paleogene stratigraphic sequences
McKenzie, D., 1978. Some remarks on the development of sedimentary and the early Andean orogenic events in the Ecuadorian Oriente Basin.
basins. Earth Planet. Sci. Lett. 40, 25–32. In: Third ISAG. St-Malo, France, 17–19, September/1996, Extended
McLaughlin Jr., D.H., 1972. Evaporite deposits of Bogotá Area, abstracts, pp. 469–472.
Cordillera Oriental, Colombia. A.A.P.G. Bull. 56 (11), 2240–2259. Rodrı́guez, C., Rojas, R., 1985. Estratigrafı́a y tectónica de la serie
Meschede, M., Frisch, W., 1998. A plate tectonic model the Mesozoic and InfraCretácica en los alrededores de San Félix, Cordillera Central de
Early Cenozoic history of the Caribbean Plate. Tectonophysics 296 (3– Colombia. In: Etayo-Serna, F., Laverde-Montaño, F. (Eds.), Proyecto
4), 269–291. Cretácico, contribuciones. Ingeominas Publ. Esp. 16, Bogotá, 21p.,
Mojica, J., Kammer, A., Ujueta, G., 1996. El Jurásico del sector norocci- Chapter XXI.
dental de Suramerica y guı́a de la excursión al Valle Superior del Roeder, D., Chamberlain, R.L., 1995. Eastern Cordillera of Colombia:
Magdalena (Nov. 1–4/95), Regiones de Payandé y Prado. Departamen- Jurassic-Neogene crustal evolution. In: Tankard, A.J., Suárez R., S.,
to del Tolima, Colombia. Geol. Colombiana, Bogotá, 21p. Welsnik, H.J. (Eds.), Petroleum Basins of South America. A.A.P.G.
Moreno, J.M., 1990. Stratigraphy of the Lower Cretaceous units central Mem. 62, 633–645.
part Eastern Cordillera, Colombia. In: 13th International Sedimento- Rolón, L.F., Carrero, M.M., 1995. Análisis estratigráfico de la sección
logical Congress, Nothingham, August 22–31, 1990. Abstract. Cretácica aflorante al oriente del anticlinal de Los Cobardes entre los
Moreno, J.M., 1991. Provenance of the Lower Cretaceous sedimentary Municipios de Guadalupe-Chima-Contratación. Departamento de
sequences, central part, Eastern Cordillera, Colombia. Rev. Acad. Col. Santander. Tesis pregrado Geologı́a, Univ. Nacional de Colombia,
Cienc. Exact. Fis. y Nat. 18 (69), 159–173. Bogotá, 80p.
L.F. Sarmiento-Rojas et al. / Journal of South American Earth Sciences 21 (2006) 383–411 411

Ross, M.I., Scotese, C.R., 1988. A hierarchical tectonic model of the Gulf Toussaint, J.F., Restrepo, J.J., 1994. The Colombian Andes during
of Mexico and Caribbean region. Tectonophysics 155, 139–168. Cretaceous times. In: Salfity, J.A. (Ed.), Cretaceous tectonics of the
Royden, L.H., 1993a. The tectonic expression of slab pull at continental Andes. Views Brunwick, Germany, pp. 61–100.
convergent boundaries. Tectonics 12, 303–325. Tron, V., Brun, J.P., 1991. Experiments on oblique rifting in brittle-ductile
Royden, L.H., 1993b. Evolution of retreating subduction boundaries systems. Tectonophysics 188, 71–84.
formed during continental collision. Tectonics 12, 629–683. Ulloa, C., Rodrı́guez, E., 1976. Geologı́a del cuadrángulo K-12. Guate-
Royden, L., Keen, C.E., 1980. Rifting process and thermal evolution of que. Bol. Geol., Ingeominas, Bogotá 22 (1), 4–55.
the continental margin of eastern Canada determined from subsidence Van der Beek, P., 1995. Tectonic evolution of continental rifts, inferences
curves. Earth Planet. Sci. Lett. 51, 343–361. from numerical modeling and fission track thermochronology. Ph.D.
Rubiano, J.L., 1989. Petrography and stratigraphy of the Villeta Group, Thesis, Vrije Universitteit, Amsterdam, 232p.
Codillera Oriental, Colombia, South America. M.Sc. Thesis, Univ. Van der Hamen, T.H., 1961. Late Cretaceous and Tertiary stratigraphy
South Carolina, Columbia, SC, 96p. and tectogenesis of the Colombian Andes. Geologie en Mijnbouw 40,
Sarmiento, L.F., 1989. Stratigraphy of the Cordillera Oriental west of 181–188.
Bogotá, Colombia. M.Sc. Thesis University of South Carolina, Van der Wiel, A.M., 1991. Uplift and volcanism of the SE Colmbian
Columbia, SC, 102p. Andes in relation to Neogene sedimentation of the Upper Magdlena
Sarmiento, L.F., 2001. Mesozoic rifting and Cenozoic basin inversion Valley. Ph.D. Thesis, Wageningen, 208p.
history of the Eastern Cordillera, Colombian Andes. Inferences from Van Wees, J.D., Stephenson, R.A., Stovba, S.M., Shymanovskyi, V.A.,
tectonic models. Ph.D. Thesis Vrije Universiteit, Amsterdam, The 1996b. Tectonic variation in the Dnieper-Donets Basin from auto-
Netherlands, 297p. mated modeling of backstripped subsidence curves. Tectonophysics
Sclater, J.G., Christie, P.A.F., 1980. Continental stretching; an explana- 268 (1-4), 257–280.
tion of the post mid-Cretaceous subsidence of the central North Sea Vergara, L., Prössl, K.F., 1994. Dating the Yavı́ Formation (Aptian
Basin. J.G.R. 85, 3711–3739. Upper Magdalena Valley, Colombia). In: Etayo Serna, F. (Ed.),
Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Estudios geológicos del Valle Superior del Magdalena. Univ. Nacional
Arispe, O., Néraudeau, D., Cardenas, J., Rosas, S., Jiménez, N., 2002. de Colombia, Ecopetrol, Bogotá, 14p., Chapter XVIII.
Late Permian-Middle Jurassic lithospheric thinning in Peru and Vesga, C.J., Barrero, D., 1978. Edades K/Ar en rocas ı́gneas y
Bolivia, and its bearing on Andean-age tectonics. Tectonophysics metamórficas de la Cordillera Central de Colombia y su implicación
345, 153–181. geológica. II Congreso Colombiano de Geologı́a, Resúmenes,
Shagam, R., 1975. The Northern termination of the Andes. In: Nairn, Bogota.
A.E.M., Stehli, H. (Eds.), The Ocean Basins and Margins, vol. 3. The Villamil, T., 1993. Relative sea level, chronology, and a new sequence
Gulf of Mexico and the Caribbean. Plenum Press, New York and stratigraphy model for distal offshore facies, Albian to Santonian,
London, pp. 325–420, Chapter 9. Colombia. In: Pindell, J.A., Drake, C.D. (Eds.), Mesozoic-Cenozoic
Smith, G.A., Landis, C.A., 1995. Intra-arc basins. In: Busby, C.J., Stratigraphy and Tectonic Evolution of the Caribbean Region/
Ingersoll, R.V. (Eds.), Tectonics of sedimentary basins. Blackwell Northern South America: Implications for Eustasy from Exposed
Science, pp. 263–298. Sections of a Cretaceous-Eocene Passive Margin Setting. Geol. Soc.
Steckler, M.S., Watts, A.B., 1978. Subsidence of the Atlantic-type Amer. Memoir, paper C-8.
continental margin off New York. Earth Planet. Sci. Lett. 41, 1–13. Villamil, T., 1994. High-resolultion stratigraphy, chronology and relaitve
Toro, G., Popeau, G., Hermelin, M., Schwabe, E., 1999. Chronology of sea level of the Albiian-Santonian (Cretaceous) of Colombia. Ph.D.
the volcanic activity and regional thermal events: a contribution from Thesis Univ. Of Colorado at Boulder, 446p.
the tephrochronology in the north of the Central Cordillera, Colom- Villamil, T., Arango, C., 1998. Integrated stratigraphy of latest Cenoma-
bia. In: Fourth ISAG. Geottingen, Germany, 4–6 October/1999, nian and early Turonian facies of Colombia. Paleogeographic evolu-
Extended Abstracts, pp. 761–763. tion and non-glacial eustasy, Northern South America. SEPM Spec.
Toussaint, J.F., 1995a. Hipótesis sobre el marco geodinámico de Colombia Publ. 58, 129–159.
durante el Mesozóico temprano, Contribution to IGCP 322 Jurassic Watts, A.B., Karner, G.D., Steckler, M.S., 1982. Lithospheric flexure and
events in South America. Geol. Colombiana, Bogotá 20, 150–155. the evolution of sedimentary basins. Phil. Trans. Roy. Soc. London
Toussaint, J.F., 1995b. Evolución geológica de Colombia 2. Triásico 305, 249–281.
Jurásico. Contribución al IGCP 322 ‘‘Correlation of Jurassic events in Wilson, M., 1993. Magmatism and the geodynamics of basin formation.
South America’’ International Geological Correlation Programme Sediment. Geol. 86, 5–29.
Unesco IUGS. Univ. Nacional de Colombia. Medellı́n, 94p. Ziegler, P.A., 1994. Geodynamic processes governing development of
Toussaint, J.F., Restrepo, J.J., 1989. Acresiones sucesivas en Colombia; rifted basins. In: Roure, F., Ellouz, N., Shein, V.S., Skvortsov, I.
un nuevo modelo de evolución geológica. V Congr. Colomb. Geol., (Eds.), Geodynamic Evolution of Sedimentary Basins. Institut Franç-
Bucaramanga I, 127–146. ais du Pétrole, Technip, Paris, pp. 16–97.

You might also like