Sei sulla pagina 1di 23

Syntectonic Cenozoic sedimentation in the northern middle Magdalena

Valley Basin of Colombia and implications for exhumation of the


Northern Andes

Elías Gómez†
Teresa E. Jordan
Richard W. Allmendinger
Department of Earth and Atmospheric Sciences, Cornell University, Ithaca, New York 14853, USA
Kerry Hegarty
Geotrack International, 37 Melville Road, Brunswick West, Victoria 3055, Australia
Shari Kelley
New Mexico Tech, Department of Earth and Environmental Sciences, Socorro, New Mexico 87801, USA

ABSTRACT INTRODUCTION mity (MMVU). While the Central Cordillera


front retreated in the middle Eocene to Recent,
Development of the Colombian Middle This paper describes the Cenozoic stra- Eastern Cordillera diachronous deformation
Magdalena Valley Basin (MMVB) was tigraphy and structure of the Colombian generated accommodation in the basin. Prior
determined by late Cretaceous-early Eocene Middle Magdalena Valley Basin (MMVB) to this work, the attributes of the MMVB were
uplift of the Central Cordillera to the west, and addresses the chronologies of deformation interpreted as linked to Central Cordillera uplift
and subsequent transferal of deformation of the bounding Central and Eastern Andean throughout most of the Cenozoic. The onset of
to the Eastern Cordillera to the east. These Cordilleras (Fig. 1). The MMVB is a part of uplift of the Eastern Cordillera proposed in this
phases are separated in the tectono-strati- a foreland basin system of major geologic and paper is also substantially earlier than that envi-
graphic record by a major unconformity, intellectual importance. In terms of volume and sioned by previous works. The dominant inter-
the Middle Magdalena Valley unconformity longevity, the Colombian foreland basin system, pretation was that the MMVB was the proximal
(MMVU). Paleocene coastal to alluvial facies consisting of the MMVB and its more distal part of a Cenozoic foreland basin coupled to the
underneath the MMVU were deposited in a neighbor, the Llanos Foothills basin, accom- Central Cordillera, which was detached from
foreland basin coupled to Central Cordillera panies the Northern Andes for over 400 km the distal Llanos Basin due to late Miocene-
kilometer-scale uplift. The middle Eocene to along strike, persisted throughout the Cenozoic, Pliocene Eastern Cordillera uplift (Schamel,
Neogene continental strata that onlap the and accumulated up to 7000 m of strata. This 1991; Dengo and Covey, 1993).
MMVU document transformation of the foreland-basin system affords an opportunity In recent years, much of the focus on the
MMVB into an interior basin due to Eastern to learn about continental deformation and relations of plate history to the tectonics of the
Cordillera deformation recorded by growth genesis of unconformities through exploration Colombian Andes has been related to Caribbean
strata in seismic lines and changing prov- of its variations from standard foreland basin tectonics (e.g., Pindell et al., 1998). The recon-
enance and paleoflow patterns. Exhumation behavior. The Colombian foreland basin system struction of deformation of the Central and
histories of the MMVB bounding ranges are initially overlapped a Mesozoic rift province, Eastern Cordilleras, described herein, informs
further constrained by apatite-fission-track and subsequent inversion of the rift system in a comparison of the evolution of the Northern
and vitrinite reflectance thermochronology the Eastern Cordillera strongly imprinted the Andes, influenced by Caribbean tectonics, to
and volcanic ash chronology. This study nature of the foreland basin. the Central and Southern Andes, where tectonic
indicates that the MMVB is fundamentally A large reserve of information about Andean activity is more simply related to subduction of
related to evolution of the entire Andean tectonic history rests in the MMVB due to its the Nazca plate. Independently, knowledge has
margin of South America. location between the two largest Andean ranges grown of the chronology and nature of tectonic
of northern South America. Syntectonic strata evolution of the Central and Southern Andes, in
Keywords: stratigraphy, tectonics, thermal and unconformities record Paleocene-early large part through investigation of basin evolu-
history, geochronology, Colombia. Eocene propagation of Central Cordillera defor- tion and improved chronologic information
mation followed by erosional retreat of this (e.g., Mpodozis and Ramos, 1990; Jordan et al.,

range. The resultant time-transgressive pedi- 1995; Horton et al., 2001). Improved knowledge
Present address: Shell International Exploration
and Production Inc., E&P Solutions, 200 North
ment surface at the base of onlapping middle of the stages and timing of uplift of the Colom-
Dairy Ashford, Houston, Texas 77079, USA; e-mail: Eocene-Neogene strata is a major unconfor- bian Andean Ranges, derived from MMVB
elias.gomez@shell.com. mity, the Middle Magdalena Valley unconfor- history, assists in distinguishing whether the

GSA Bulletin; May/June 2005; v. 117; no. 5/6; p. 547–569; doi: 10.1130/B25454.1; 15 figures; 3 tables; Data Repository item 2005076.

For permission to copy, contact editing@geosociety.org


© 2005 Geological Society of America 547
548
B

Panam

Venezuela

Colombia
GÓMEZ et al.

South
America
Ecuador
Per

Geological Society of America Bulletin, May/June 2005


0 300 km

Figure 1. (A) Shaded relief map of Colombia and nearby regions illus-
trating the general tectonic and physiographic setting of the Middle
Magdalena Valley Basin (MMVB) within white polygon, and surround-
ing Andean ranges in northwestern South America. (B) Geologic map
of the MMVB and adjacent mountain ranges and basins showing the
locations of structural cross sections and main structures discussed in
this paper. Map sources: Geotec, 1988; Schamel, 1991.
SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

Central and Eastern Cordilleras of Colombia are


highly similar to the Central Andes but slightly
overprinted by Caribbean or collisional influ-
ences or instead are markedly different.
Traditionally, a distinction has been made
between the Northern Andes and the Central
Andes in relation to the role of terrane accretion.
Colombia’s oceanic Western Cordillera was
progressively accreted from Cretaceous through
Eocene time (Barrero, 1979; Etayo-Serna et al.,
1983), whereas the Panamá-Baudó arc accreted
during the middle Miocene to Pliocene (Duque-
Caro, 1990; Dengo and Covey, 1993; Fig. 1A).
A second category of distinction between the
Central Andes and Northern Andes is caused by
extensive tectonic erosion of the Central Andean
margin (removal of the forearc lithosphere by
subduction), which controlled the preservation
of the western margin and may have caused
significant eastward migration of the Central
Andean uplift. Because of these distinctions, we
emphasize a comparison of the tempo of tectonic
evolution of the Northern and Central Andes. We
use the northern Chile–southern Bolivia–north-
ern Argentine Andes (20–24°S) as the Central
Andean reference zone. Major steps in Central
Andean evolution that merit comparison are
Jurassic-early Cretaceous backarc extension, late
Cretaceous-Paleocene compression, a Paleo-
cene-Eocene interval of broad erosion and local-
ized deformation, an Oligocene-early Miocene
interval of reorganization of deformation, and
lastly, a major surge in the degree of continental
shortening early in the Miocene, which led to the
development of the modern Andean volcanic arc
and thrust systems.
From a resource perspective, the Colombian
foreland basin system has been one of the most
prolific petroleum areas of the entire Andean
belt. This paper provides data relevant to the
MMVB main oil reservoirs and traps, which
are of Cenozoic age. Maturity trends of marine
source rocks, found within the underlying Cre-
taceous sequence, are also partially explained
by Cenozoic tectonics and sedimentary burial.

DATA, METHODS, AND


ORGANIZATION OF THIS PAPER

Surface and subsurface descriptions of stra-


tigraphy and structure form the basis of our
studies. Surface stratigraphic columns provided
facies, paleocurrent, and petrographic data
(Figs. 2 and 3). Regional distributions of sedi-
mentary units and unconformities were mapped
Figure 2. Geologic map of the northern and central MMVB showing the locations of struc- in the subsurface in a grid of 2800 km of 2-D
tural and seismic sections as well as stratigraphic columns and places mentioned in this reflection seismic lines and 99 well logs with
paper. See location in Figure 1B. Map sources: Servicio Geológico Nacional (1966, 1967), formational tops provided by several oil com-
Ward et al. (1977). Locations of apatite-fission-track samples and values of vitrinite reflec- panies. Selected seismic lines were converted to
tance (Ro), discussed in text, are also shown. depth using velocity models derived from well

Geological Society of America Bulletin, May/June 2005 549


GÓMEZ et al.

Figure 3. Composite stratigraphic column illustrating the stratigraphic subdivisions (after Morales et al., 1958) and main lithologic character-
istics of the MMVB Cenozoic deposits. Field descriptions were originally conducted at the decimeter scale; thicknesses were measured with a
1.5-m-long Jacob staff and tape and compass. The stratigraphic positions of petrographic samples are also illustrated; samples with an asterisk
were also used for apatite-fission-track thermal history analyses. Vitrinite reflectance samples of the Lisama, La Paz, and Esmeraldas forma-
tions are underlined. The positions of zircon-fission-track dated reworked tuffs appear in bold italics. See Figure 2 for location of measured
sections; outcrops of the Los Corros and La Cira fossil horizons were studied at the localities of Uribe-Uribe and Oponcito River, respectively.
Refer to Table 2 and Figure 5 for detailed facies descriptions of the Lisama, La Paz, Esmeraldas, Colorado, and Real units.

550 Geological Society of America Bulletin, May/June 2005


SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

data and stacking velocities. Uplift and erosion and Delaloye, 1982; Bayona et al., 1994). A pattern may reflect the distribution of Mesozoic
episodes of the MMVB bounding ranges are marine transgression of the Colombian territory normal faults, whose reactivation as reverse
also constrained by structural cross sections and during the Cretaceous led to a transgressive- faults is interpreted to have driven Eastern Cor-
growth strata displayed in seismic data from the regressive megasequence mostly composed of dillera uplift (Schamel, 1991). The master faults
Eastern Cordillera foothills. Thermal histories shales, limestones, and chert beds (Macellari, of this system are known as the La Salina fault
derived from apatite-fission-track analyses and 1988), which reaches a maximum thickness of system. Trailing anticlines are fault-bend folds
vitrinite reflectance studies provide additional 4 km in the Eastern Cordillera foothills along cored by Mesozoic rocks, whereas hanging-wall
information on episodes of uplift and sedimen- the northern MMVB. synclines carried above detachment surfaces are
tary burial. We also report two fission-track ages Late Cretaceous to Eocene oblique accre- composed mostly of Tertiary rocks. Hanging
from upper Miocene volcanoclastic beds. tion of the Western Cordillera (Fig. 1A) caused wall fault-propagation anticlines, sometimes
After describing the MMVB geologic setting, northward propagation of uplift of the Central associated with blind thrusts, constitute the
we evaluate the depositional ages and describe Cordillera (Campbell, 1968). Tectonic inversion westernmost structures of the Eastern Cordillera
the facies and composition of Cenozoic units. of Mesozoic grabens in the Eastern Cordillera foothills along the MMVB.
We continue with descriptions of the subsurface area also initiated at that time and continued
stratigraphic and structural geometry of the throughout the Cenozoic (Julivert, 1963; Gómez, AGE, OUTCROP STRATIGRAPHY, AND
MMVB and present relevant thermal-history 2001). But the most intense pulse of Eastern Cor- PETROGRAPHY OF CENOZOIC ROCKS
information. Regional and local data sets are dillera uplift has been attributed to accretion of
then synthesized into palinspastic reconstruc- the Panamá-Baudó arc, which started at 12.9 Ma In this section, we describe the ages, facies,
tions of MMVB evolution. Finally we compare (Duque-Caro, 1990; Dengo and Covey, 1993). and compositions of the MMVB Tertiary rocks
the mechanisms of tectonic evolution of the The structural configuration of the Cen- as derived from surface data. The most complete
Andean foreland basin system of Colombia with tral Cordillera is not well understood. It may and best exposed Paleocene-lower Oligocene
the Central Andes of South America. involve a combination of strike-slip and reverse rock record (Lisama, La Paz, and Esmeraldas
faulting such that the Central Cordillera may formations) occurs in the eastern flank of the
GEOLOGIC SETTING represent a crustal-scale, thick-skinned, positive Nuevo Mundo syncline along the Sogamoso
flower structure (D. Barrero, personal commun., River; outcrops of the overlying Oligocene-
The MMVB overlies continental-affinity 1989, 1999). A significant component of com- Miocene rocks (Mugrosa, Colorado, and Real
basement of Proterozoic to Paleozoic age, which pressional deformation must have occurred to units) are found in the La Cira-Infantas oil field
also constitutes the crystalline core of the Cen- generate important uplift and exposure of pre- area along the Pan-American road (Figs. 2 and
tral and Eastern Andean Cordilleras (Fig. 1A). Mesozoic basement, which reaches altitudes 3; Morales et al., 1958; Ramón, 1998).
To the west of the Romeral fault system, along of 3500 m in places where it is not overlain by
the western flank of the Central Cordillera, are Pliocene volcanic edifices. The age of the major Chronology of MMVB Cenozoic Fill
oceanic terranes, which were accreted to South phase of deformation of the Central Cordillera
America through several collisions from the at the latitude of the MMVB is pre-middle The most important stratigraphic bound-
early Cretaceous through the Cenozoic (Bar- Eocene, as constrained by relations across the ary within the MMVB Cenozoic record is the
rero, 1979; Etayo-Serna et al., 1983). To the east MMVU described below. MMVU, between Paleocene and middle Eocene
of the Romeral fault (Fig. 1A), granulite-grade The Eastern Cordillera is flanked by fold- units in Figure 3. We defer the description of
metamorphic basement in the Eastern Cordillera and-thrust belts with opposite vergence, which this surface and other Cenozoic unconformities
is overlain by Paleozoic low-grade metasedi- overthrust the MMVB and Llanos Basin to the to a later section in this paper. Here, we review
mentary rocks (Restrepo, 1995). More extensive west and east, respectively (Fig. 1B). The NE the age assignments of Cenozoic units as a
exposures of crystalline basement are found in structural trend of the Eastern Cordillera changes first step to understanding the evolution of the
the Central Cordillera; low- to medium-pressure to NNW north of 7°N, and it also splits into two MMVB (Fig. 4). The palynological assignment
metasedimentary rocks are dominant in this Andean branches, the Santander Massif and of the Lisama Formation to the Paleocene (Van
range (Etayo-Serna et al., 1983). Mesozoic and Perijá Range to the NW, and the Mérida Andes der Hammen and Garcia, 1966) has not been
Cenozoic calc-alkaline plutons intrude the meta- to the NE. The Bucaramanga Fault constitutes disputed. In contrast, various authors working
morphic complexes of the Eastern and Central the western boundary of the Santander and Santa with different fossil groups reached conflicting
Cordilleras and the oceanic terranes to the west Marta Massifs and is a NNW-trending left-lat- interpretations of the ages of the middle Eocene
of the Romeral fault (Aspden et al., 1987). eral strike-slip fault with a major component of to lower middle Miocene La Paz, Esmeraldas,
Mesozoic rifting characterized the MMVB west-verging, reverse movement (Fig. 1; Camp- Mugrosa, and Colorado formations. Only the
and the Eastern Cordillera areas (Etayo-Serna bell, 1968). Estimates of the amount of sinistral upper Miocene Real Group and the Pliocene
et al., 1983). Triassic to Jurassic synrift red displacement of the Bucaramanga fault range Mesa Group have been dated by geochronology
beds locally reach 5000 m in thickness and are between 100 and 115 km; offset features include of interbedded volcanic rocks.
exposed in the cores of the Eastern Cordillera crystalline and Mesozoic stratigraphic units of The Eocene-Miocene stratigraphy was based
Arcabuco and Los Cobardes anticlines and in the Central Cordillera relative to the Santa Marta on identification of the Los Corros, Mugrosa, and
the San Lucas Range of the Central Cordil- Massif and the Ranchería Basin relative to the La Cira fossil horizons located at the tops of the
lera (Fig. 1B; Cediel, 1968; Bogotá and Aluja, northern MMVB (Fig. 1B; Campbell, 1968; Esmeraldas, Mugrosa, and Colorado formations,
1981). Mesozoic extension in these localities Forero, 1974; Pindell et al., 1998). respectively (Fig. 3). These horizons consist of
is recorded by unconformity-bounded units, In map view, anticline-syncline pairs arranged thin layers of packstones of bivalves and gastro-
rotated-block morphology, spatially variable in an eastward stepping pattern from south to pods within muddy intervals that are up to 15 m
thicknesses, bimodal subaerial volcanogenic north characterize the Eastern Cordillera thrust- (Los Corros), 8 m (Mugrosa) and 106 m thick
strata and mafic intrusives (Julivert, 1958; Fabre and-fold belt along the MMVB (Fig. 1B) This (La Cira; Wheeler in Pilsbry and Olsson, 1935).

Geological Society of America Bulletin, May/June 2005 551


GÓMEZ et al.

The Los Corros and La Cira are composed of Cenozoic Facies, Sedimentary Environments,
fresh-water mollusks and a few slightly brack- and Paleoflow
ish-water elements; the Mugrosa Fossil horizon
is composed exclusively of fresh-water mollusks Decimeter-scale descriptions and measure-
(Pilsbry and Olsson, 1935; Nuttal, 1990). The La ments of lithofacies, facies associations, and
Cira horizon can be traced toward the south as paleoflow of the MMVB Cenozoic units were
far as the boundary with the upper Magdalena carried out in the Eastern flank of the Nuevo
Valley Basin (Fig. 1B; Nuttal, 1990), whereas Mundo syncline and in the La Cira-Infantas
the Los Corros and Mugrosa fossil horizons oil field area (Figs. 2, 3, and 5; Table 2). The
are restricted to the central and northern parts of Paleocene Lisama Formation overlies the Maas-
the MMVB. Cenozoic age assignments derived trichtian shallow marine Umir Formation with a
from the fossil horizons are imprecise. Pilsbry transitional contact and records regressive sedi-
and Olsson (1935) tentatively assigned a late mentation in deltaic and alluvial plains (Table 2,
Eocene age for Los Corros, a middle Oligocene Fig. 5; Ramírez, 1988; Gómez, 2001). The La
age for the Mugrosa, and a late Oligocene-early Paz Formation mudstones and sandstones above
Miocene age for La Cira Horizon. the MMVU were deposited in fluvial settings and
Somewhat different palynological ages have indicate complete transformation of the MMVB
been assigned to the La Paz Formation in the into a continental basin (Ramírez, 1988; Suárez,
same section in the eastern flank of the Nuevo 1997). Fossiliferous packstones and mudstones
Mundo syncline (Figs. 1B and 2). This unit is of the Los Corros, Mugrosa, and La Cira hori- Figure 4. Synthesis of age assignments for
late Eocene according to Ramírez (1988) and zons (Fig. 3, not shown in Fig. 5) record the the MMVB Cenozoic units. Vertical bars
middle Eocene according to Rueda (1996, in establishment of mollusk communities in well- describe the maximum age range of each
Olaya, 1997). For younger strata, Nuttal (1990) oxygenated ephemeral lakes or backswamps in unit permissible by data from the eastern
reviewed the work of palynologists who based the MMVB floodplain. Slightly brackish mol- flank of the Nuevo Mundo syncline (Lisama,
the age assignments of pollen zones on plank- lusks of the Los Corros and La Cira horizons La Paz, and Esmeraldas formations) and the
tonic foraminifera zones (e.g., Hopping, 1967) may relate to minor marine influence. La Cira-Infantas area (Mugrosa, Colorado,
and who applied those ages to palynological Two distinct units compose the La Paz Forma- Real, and Mesa units). Uncertainties of the
evidence associated with the MMVB fossil tion in the eastern flank of the Nuevo Mundo paleontologic age assignments of the bound-
horizons. He stated that the Los Corros horizon syncline (Figs. 3 and 5). Bioturbated silty mud- aries of Eocene–early Miocene units may be
is more likely to be early Oligocene rather than stones predominate in the lower La Paz, whereas as great as 5–10 m.y. Ages of the Real and
Eocene; the Mugrosa is late Oligocene, and the upper La Paz is a multistoried stack of Mesa Groups are constrained by zircon-fis-
the La Cira is middle Miocene (Nuttal, 1990). cross-bedded sandstones. In general, the La Paz sion-track and K/Ar dating, respectively.
Other MMVB palynologic determinations Formation at this locality displays an upward- The superimposed thicker bars represent
(Ramírez, 1988; Rueda, 1996 in Suárez, 1997 coarsening trend of facies, which is followed our preferred interpretation of depositional
and Olaya, 1997) place the Los Corros in the by an upward-fining trend that continues into ages. See text for discussion and sources of
late Eocene, the Mugrosa horizon in the early the Esmeraldas Formation. This trend records data and further discussion. The time gap
Miocene, and the La Cira horizon in the early to autocyclic wandering of fluvial channels across associated with the MMVU time-transgres-
middle Miocene. the MMVB floodplains (Gómez, 2001). Axes sive surface is the shortest in the Eastern
The age of the Real Group is constrained by of trough cross-bedding indicate a predominant flank of the Nuevo Mundo syncline; see
palynological and foraminiferal comparisons NE orientation of paleoflow for the La Paz and descriptions of subsurface structure and
and by radiometric dating. Pollen found at the Esmeraldas formations in the eastern flank of the stratigraphy of the MMVB in the text for
base of the Real Group, just above the La Cira Nuevo Mundo syncline. further information on time and geographic
horizon, lies within the lower part of the forami- The Mugrosa Formation conformably overlies distributions of the MMVU. Geologic time
niferal zone Globorotalia fohsi fohsi (Hopping, the Esmeraldas Formation in the La Cira-Infantas scale after Berggren et al. (1995).
1967; Nuttal, 1990), whose time span is brack-
eted between 12.7 Ma and 12.5 Ma (Berggren et
al., 1995). We obtained two zircon-fission-track
ages from reworked tuffs in the upper part of TABLE 1. ZIRCON-FISSION-TRACK AGES FOR THE REAL GROUP
the Real Group (samples PA-V4c and PA-V13;
Sample Number of ρs × 105 ρi × 106 ρd × 105 Central age Chi2 Uranium content
Fig. 3, Table 1). They gave ages of 7.0 ± 1.2 and number grains (tracks/cm2) (tracks/cm2) (tracks/cm2) (Ma ± 2 S.E.) (%) (ppm)
6.2 ± 0.8 Ma, which are interpreted as related
PA-V4c 19 6.28 11.65 2.84 7.0 ± 1.2 50 513
to volcanic activity that was nearly simultane- (190) (1763) (4010)
ous with deposition. Thus, the maximum age PA-V13 20 7.35 15.66 2.97 6.2 ± 0.8 97 685
of deposition of the Real group is constrained (489) (5212) (4010)
to the late middle to late Miocene time interval Notes: ρs—spontaneous track density. ρi—induced track density (reported induced track density is
12.7 Ma to 6.2 ± 0.8 Ma (Serravallian to Mes- twice the measured density); number in parentheses is the number of tracks counted for ages and
sinian). The Mesa Group in its type area, the fluence calibration or the number of tracks measured for lengths. ρd—track density in muscovite detector
covering CN-5 (10 ppm); reported value determined from interpolation of values for detectors covering
southernmost MMVB, is a volcanoclastic unit standards at the top and bottom of the reactor packages (fluence gradient correction). S.E.—standard
dated as Pliocene, based on K/Ar dating of pum- error. Chi2—Chi-squared probability. λf = 1.551 × 10-10yr–1, g = 0.5. Zeta = 492 ± 50 for apatite, 449 ± 49
for zircon.
ice levels (Thouret, 1989).

552 Geological Society of America Bulletin, May/June 2005


SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

TABLE 2. LITHOFACIES OF THE CENOZOIC SEDIMENTARY FILL OF THE NORTHERN MMVB


Facies Description Interpretation
code
Gm Meter-scale thick layers of horizontally bedded, clast supported, subrounded, pebble Bedload sheets deposited in deepest parts of gravelly fluvial
conglomerates. Beds display irregular bases and frequent pebble imbrication. Gm channels (thicker beds), lag deposits at the bases of
also appears as thin lenses at the bases of St/Sla sandstones. sandy channels (thin lenses) (Allen, 1983; Miall, 1988).
Gt Meter-scale thick beds of clast supported, subrounded to rounded, pebble Upper-flow regime migration of linguoid gravel bars (Allen,
conglomerates. Internal structure consists of decimeter-thick sets of trough cross- 1983; Miall, 1988).
bedding. Bases of beds are sharp and irregular.
St Fining upward, decimeter- to meter-scale thick beds of fine- to medium-grained Scouring of channels followed by migration of lower-flow
sandstones with internal trough cross-bedding. Beds are elongate or lens-shaped regime sand dunes along the bottom (Miall, 1988).
and have irregular bases.
Sla Similar to St but conspicuous inclined surfaces of discontinuity (lateral accretion Fluvial point-bar deposits. Lateral accretion surfaces bound
surfaces) transect the sandstone beds from base to top. high-stage sedimentary increments (Miall, 1992).
Sr Centimeter- to decimeter-thick beds of very fine and fine-grained sandstones with Lower-flow regime sedimentation in fluvial channels, natural
ripple cross-lamination. Beds are elongate or lens-shaped and display undulated levees, and crevasse splays (Miall, 1988, 1992).
bases. Dessication cracks and ferruginous crusts are common at the top.
Sm Centimeter- to decimeter-thick beds of very fine and fine-grained sandstones. Beds Falling-stage channel deposits and bioturbated crevasse
display intense burrowing and root bioturbation. splay deposits (Miall, 1988, 1992).
Fl Millimeter- to centimeter-thick interlaminations of mudstone, siltstones, and very fine Passive channel fill or overbank deposits (Miall, 1988,
grained sandstones. 1992).
Fm1 Meter-scale thick beds of variegated mudstones with varying degrees of bioturbation. Flood basin deposits and paleosoil formation (Retallack,
1988).
Fm2 Meter-scale thick beds of light gray, siliceous, silty mudstones, with abundant oxidized, Flood basin deposits. Development of well-drained soils
vertical root casts. (Retallack, 1988).
Fcf Decimeter-scale thick beds of dark gray mudstone containing fresh and brackish water Back swamp, lake sedimentation (Miall, 1988, 1992).
fossil mollusks.
Pcf Centimeter- to decimeter-scale thick beds of packstones of mollusk remains. Back swamp, lake sedimentation (Miall, 1988, 1992).
Fco Centimeter- to decimeter-scale thick beds of dark gray, pyritic, carbonaceous Poorly drained swamp deposits (Miall, 1988, 1992).
mudstones.
C Millimeter- to centimeter-thick coal seams. Poorly drained swamp deposits (Miall, 1988, 1992).
Vr Reworked andesitic tuffs. Channel reworking of volcanic ash and lapilli deposits.
Note: Gravel, sandstone, and mudstone facies codes taken from or modified after Miall (1988).

oil field locality. Its lower part, known opera- MMVB areas to the west of the Eastern Cordil- (Mesa, 1995; Fig. 6A). The components of the
tionally as the “C sandstones” in the subsurface lera thrust-and-fold belt. Cenozoic sandstones are grains of polycrystal-
of the La Cira-Infantas field, does not crop out line and monocrystalline quartz, metamorphic
in this area (Dickey, 1992; Gutiérrez, 2001). The Petrography of the MMVB Cenozoic fragments (schist, micaceous gneiss, and phyl-
exposed upper Mugrosa Formation, equivalent Sandstones and Conglomerates lite), sedimentary clasts (chert and mudstone),
to the subsurface “B sandstones” of the La Cira- feldspar grains, and scarce igneous fragments
Infantas oil field (M. Gutiérrez, 2000, personal A subset of 11 sandstone samples from the (Fig. 6B). Metamorphic grains are the most
commun.), display well-preserved sandstone- Lisama, La Paz, Esmeraldas, and Real units abundant lithic components of the Lisama For-
mudstone point-bar sequences, which reflect was selected for thin-section analyses, which mation, below the MMVU. Above the MMVU,
high rates of accommodation relative to sedi- involved identifying 300 fine- to medium- the percentage of sedimentary grains increases
ment supply (Fig. 5; Ramón, 1998). Thick beds grained sand grains (Fig. 6). Half of each thin abruptly, largely replacing the metamorphic
of epiclastic volcanic material appear in the section was stained with sodium cobaltinitrite rock fragments. Sedimentary grains of the
upper half of the Real Group section. Zircon-fis- to identify potassium feldspar. Mesa (1995) lower La Paz are composed predominantly of
sion-track ages, reported before, were obtained studied the petrography of the Mugrosa and dark mudstones with organic-matter streaks
from two of these beds. In the La Cira-Infantas Colorado sandstones on 296 thin sections from and foraminifera casts. Potassium feldspar
area, trough cross-bedding indicates southeast- well cores in the La Cira-Infantas oil field. increases and polycrystalline quartz and sedi-
ward and northward paleoflow directions for This author identified 350–400 grains per thin mentary grains decrease upward into the upper
the upper Mugrosa Formation and Real Group, section; a subset of 148 thin sections was also La Paz. The decreasing trends in contents of
respectively. stained with sodium cobaltinitrite in order to polycrystalline quartz and sedimentary grains
The Pliocene Mesa Group rests conformably identify potassium feldspar (Mesa, personal continue upward into the sandstones of the
upon the Real Group and is reportedly com- commun., 2004). Esmeraldas Formation.
posed of massive conglomerates, cross-bedded Most of the MMVB Cenozoic sandstones Mesa’s (1995) study reveals a homogeneous
lithic sandstones, and mudstone layers; this unit are litharenites and sublitharenites with the composition of the Mugrosa and Colorado
is 575 m thick (Morales et al., 1958). Occur- exception of the Mugrosa and Colorado sand- feldspathic sandstones. Of the total feldspar
rence of the Mesa Group is restricted to the stones, which are predominantly feldspathic grains, 15%–30% of sand-size grains, the ratio

Geological Society of America Bulletin, May/June 2005 553


554
GÓMEZ et al.

Geological Society of America Bulletin, May/June 2005


Figure 5. Summary of MMVB Cenozoic facies, paleoflow, and interpretations of sedimentary environments. Refer to Table 2 for facies codes and Figure 3 for lithofacies and
symbols legend. In all cases, rose diagrams of paleoflow synthesize measurements of axes of 3D exposures of trough cross bedding. Each measurement corresponds to indi-
vidual channel sandstone bodies within multistoried stacks of sandstones (e.g. upper La Paz) or embedded within mudstones (all other units).
SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

A quartzite (40%–95%). Chert pebbles of the Real


Group contain Cretaceous foraminifera.

SUBSURFACE STRATIGRAPHY AND


STRUCTURE OF THE MMVB

Description of the Middle Magdalena


Valley Unconformity (MMVU)

The MMVU is a major tectonic and strati-


graphic boundary that truncates deformed
pre-Eocene rocks throughout the MMVB and
Eastern Cordillera foothills; a succession of rel-
atively undeformed middle Eocene to Neogene
strata onlap this surface. As illustrated by Fig-
B ures 7 and 8, the MMVU dips regionally toward
the east, with modifications near the Infantas
paleohigh (Morales et al., 1958) and small NE-
trending anticlines in the northernmost MMVB
(Cáchira arch). The NE-trending anticlines
are underlain by positive flower structures,
the probable result of strike-slip deformation
(Fig. 7A). Westward, the MMVU is continuous
with the eastern slope of the Central Cordillera
along most of the MMVB (Figs. 7C and 8), but
such continuity is locally disturbed by the Can-
tagallo fault (Fig. 7B). Surface geologic map-
ping indicates intense faulting of the Central
Cordillera crystalline rocks (e.g., Geotec Ltda,
1988). Figure 8A shows the plan-view distri-
bution of crystalline basement and Mesozoic
and Paleocene sedimentary rocks underneath
the MMVU. The subsurface area of preserved
Cretaceous rocks is larger than the area of
Paleocene deposits, which are restricted to the
Figure 6. Thin-section petrography of northern MMVB sandstones. (A) Classification of northeastern portion of the MMVB. Region-
MMVB sandstones (e.g., Folk, 1974). (B) Variations of MMVB sandstone grain composi- ally, the thickness of preserved Mesozoic and
tions through Cenozoic times. The stratigraphic positions of the samples are indicated in Paleocene rocks increases eastward toward the
the right side of these figures (see also Fig. 3). These figures also synthesize Mesa’s (1995) foothills of the Eastern Cordillera and decreases
petrographic study of the Mugrosa and Colorado formations in the La Cira-Infantas area. northward toward the Cáchira arch (Figs. 7B
See text for further discussion. A table summarizing the thin-section petrographic determi- and 7D). Angularity between units below and
nations for the Lisama, La Paz, Esmeraldas, and Real units is available in the GSA Data above the MMVU decreases toward the east
Repository (Table DR1).1 across the MMVB. In the eastern flank of the
Nuevo Mundo Syncline (Fig. 7B), the angular
relation between the Lisama beds and onlap-
ping La Paz strata is about one degree, which
of potassium feldspar to plagioclase is ~3–1. age (traces–2%) of hornblende in this section of makes the MMVU difficult to distinguish at the
An increase in sedimentary grains and decrease the Real Group. surface. However, an early to middle Eocene
in feldspar percentage differentiate the Real Conglomeratic levels are mostly restricted sedimentary hiatus at this locality is revealed by
Group from the Colorado sandstones. Wheeler to the La Paz and Colorado formations and the ages of the Lisama and La Paz Formation
(in Pilsbry and Olsson, 1935) indicated that the the Real Group. Gravel compositions were (e.g., Fig. 4).
Real sandstones are characterized by abundant determined for 120–200 clasts in representative
hornblende, which is rarely present in older 1-m2 outcrop areas. The conglomeratic beds at Distribution of Onlapping Middle Eocene
rocks. However, we only found a small percent- the base of the La Paz are made of pebbles of to Neogene Units
white quartzite (78%), siliceous siltstone and
chert (15%), and rock fragments such as phyl- The thickness of the middle Eocene to
1
GSA Data Repository item 2005076, thin-sec- lite (4%), red siltstone (1%), sandstone, and Neogene deposits above the MMVU increases
tion petrography of Cenozoic sandstones from the
northern MMVB, is available on the Web at http://
rhyolite (2%). Granules and pebbles of the Col- regionally toward the ESE, as shown by the con-
www.geosociety.org/pubs/ft2005.htm. Requests may orado Formation and Real Group are composed toured depth to the MMVU (i.e., thickness of
also be sent to editing@geosociety.org. predominantly of siliceous siltstone and white post–early Eocene strata, Figs. 8A and 8B). The

Geological Society of America Bulletin, May/June 2005 555


GÓMEZ et al.

B, C, D

Figure 7. Structural cross sections perpendicular (A, B, C), and parallel (D) to the MMVB strike; sections are based on surface and subsur-
face information. See Figures 1B and 2 for locations. (A) Structure underneath the MMVU is characterized by positive flower structures.
(B) The Infantas paleohigh was buried during sedimentation of the upper Eocene–lower Oligocene (?) Esmeraldas Formation. The Can-
tagallo fault is an early Paleogene structure that was reactivated during the Pliocene. (B)–(C) The Eastern Cordillera foothills consist of
two superimposed fold-and-thrust belts of late Oligocene to early middle Miocene (e.g., Lisama anticline) and Pliocene (Guayabito, Opón
anticlines) ages. Both generations of folds are truncated by out-of-sequence faulting (e.g., La Salina fault). (D) The northern Cáchira arch
was overlapped by the lower middle Miocene topmost beds of the Colorado Formation. Sections (A), (C), and (D) also reveal truncation of
tilted Cenozoic strata beneath the MMVB. Sections (B) and (D) show the location of apatite-fission-track samples, accompanying values of
vitrinite reflectance (Ro), and derived thermal history interpretations. See Figure 2 for stratigraphic locations of Paleogene AFTA and Ro
samples in section (B). The La Salina reverse fault is a reactivated Mesozoic normal fault as highlighted by the differences in thickness of
the Jurassic synrift rocks (e.g., Cediel, 1968; Rabe, 1977) across the fault in sections (B) and (C).

556 Geological Society of America Bulletin, May/June 2005


SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

N N

Figure 8. (A) Contoured depth to the MMVU in the subsurface of the MMVB between the Central and Eastern Cordilleras and subcrop distribu-
tion of Paleocene and older rocks beneath the MMVU, according to the seismic and well data. Contour map of the MMVU modified after Geotec-
Maxus (1993). (B) “Worms-eye view” of onlapping middle Eocene to Neogene sedimentary units above the MMVU, according to seismic and well
data. Other sources of data: Morales et al. (1958), Ramírez (1988), Olaya (1997), Suárez (1997), Ramón (1998). See text for discussion.

middle Eocene-Neogene succession reaches a MMVB informally called the Cantagallo sand- stack of conglomeratic sandstones, interpreted
maximum thickness of 7000 m in the footwall stones (Suárez, 1997). The upper La Paz mul- as an alluvial fan deposit associated with activ-
of the La Salina fault (Eastern Cordillera foot- tistoried sandstones are missing on the western ity of the Cantagallo fault (Suárez, 1997).
hills). The time gap represented by the MMVU flank of the Nuevo Mundo syncline, replaced The upper Eocene-lower Oligocene Esmer-
decreases from west to east as shown by com- by flood-plain mudstones and sandstones, aldas Formation is more broadly distributed
parison of Figures 8A and 8B. Sedimentary described by Suárez (1997) in well cores from and is everywhere younger than the La Paz. In
onlap of the MMVU began with the La Paz and the Provincia anticline (Fig. 2). The westward most of the northern MMVB, an unconformity
Esmeraldas formations (Figs. 7B, 7C, and 8B). change to finer-grained facies occurs as the La separates the La Paz and Esmeraldas units. The
The middle-upper Eocene La Paz is limited in Paz thins to the west. Description of the Can- La Paz-Esmeraldas unconformity is expressed
extent to the eastern quarter of the MMVB, with tagallo sandstones also relies on well data and on seismic data as truncated reflectors of the
a correlative accumulation in the westernmost subsurface mapping. This unit is a 600-m-thick underlying Cantagallo sandstones and onlap

Geological Society of America Bulletin, May/June 2005 557


GÓMEZ et al.

terminations of the overlying Esmeraldas reflec-


tors (Figs. 9 and 10). The Oligocene-lower mid-
dle Miocene Mugrosa and Colorado formations
extend farther to the west and north (Figs. 7C,
7D, and 8B). The Cáchira Arch was a structural
high in the northernmost MMVB throughout
the sedimentation of middle Eocene-lower Mio-
cene rocks. Only middle and upper Miocene
rocks represented by the uppermost Colorado
Formation and the Real Group are found above
Cretaceous rocks in the northernmost part of the
MMVB. The Real Group occurs throughout the
MMVB. The base of the Real Group is con-
formable in the central parts of the basin areas
but becomes an angular unconformity toward
the east, where these rocks rest upon the eroded
crests of late Oligocene-early middle Mio-
cene folds in the Eastern Cordillera foothills,
as shown below. Throughout the MMVB, a
regional unconformity truncates southeastward
dipping Neogene strata (e.g., Fig. 7; Ramírez,
1988). This unconformity coincides with the
present surface of the MMVB and joins the
MMVU (i.e., Central Cordillera slope) toward
the west.

Structure, Growth Strata, and Deformation


Chronology of the Eastern Cordillera
Foothills

Décollement surfaces and footwall ramps


of the Eastern Cordillera thrust-and-fold belt
occur at several stratigraphic levels along the
MMVB (Figs. 7B and 7C). This variability
has two causes. First, Cretaceous and Paleo-
cene rocks, mainly incompetent shales, are
unevenly distributed due to subsequent ero-
sion. Second, the MMVB Tertiary stratigraphy Figure 9. Depth-converted seismic section across the western side of the MMVB, to the west
is characterized by strong lateral and vertical of the Infantas paleohigh. See Figure 2 for location. Strata of the Cantagallo sandstones
facies changes between competent sandstones (equivalent to the La Paz Formation) onlap the MMVU. Reflectors of the Cantagallo sand-
and incompetent mudstones. stones are also truncated against a subtle unconformity that separates this unit from the
Eastern Cordillera folds displayed in Fig- overlying Esmeraldas Formation. This unconformity merges with the MMVU toward the
ure 7B formed due to slip on the La Salina fault east margin of the seismic section. Cenozoic units above the composite Cantagallo-Esmer-
system below them. From east to west, the La aldas-MMVU surface show conformable relations in this section. The Cantagallo fault was
Salina fault ramps across Cretaceous rocks and active during the deposition of the Cantagallo sandstones and Esmeraldas Formation and
has a footwall flat in the upper Cretaceous rocks, was reactivated during the Pliocene–Pleistocene, as revealed by cross-cutting relations with
which splays into two ramps across Paleogene the Real Group. See text for discussion.
units. The ramps, as preserved in the footwall,
are listric. The Lisama anticline is a hanging-
wall fault-propagation fold associated with the
westernmost leading thrust ramp, which has its Real Group and it is not folded by the underlying Growth strata of the same age are also found at
tip in the Mugrosa Formation. A minimum early Lisama anticline. The eastern flank of the Nuevo the backlimb of the Provincia anticline (10 km
middle Miocene time limit for the formation of Mundo syncline corresponds to the forelimb of to the north, Figs. 2 and 12) and the forelimb of
the Lisama anticline is set by its position under- the Los Cobardes fault-bend anticline. the Los Cobardes anticline (Fig. 10). Smaller-
neath undeformed beds of the Real Group. The The age of formation of the Lisama anticline scale units of onlapping strata are bounded by
Nuevo Mundo syncline is bounded to the west is better constrained by a wedge of divergent unconformities (i.e., growth unconformities) in
by the second splay of the major La Salina fault growth strata, equivalent in age to the upper the backlimb of the Provincia anticline (Fig. 12).
system. This splay juxtaposes a hanging wall flat Mugrosa and Colorado formations (upper Oli- The overall geometry of these synorogenic
above a footwall ramp. This is the youngest ramp gocene-lower middle Miocene), which flanks strata is consistent with sedimentary accumu-
of the thrust system, as it cuts across beds of the the backlimb of the Lisama anticline (Fig. 11). lation contemporaneous with progressive limb

558 Geological Society of America Bulletin, May/June 2005


SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

Figure 10. Depth-converted seismic section across the Nuevo Mundo syncline. See Figure 2 for location. At this location, the MMVU is a
subtle angular unconformity that separates the Lisama Formation from the onlapping La Paz Formation. Another subtle angular uncon-
formity constitutes the boundary between the La Paz Formation and the overlying Esmeraldas Formation. Strata of the middle Eocene to
lower Oligocene La Paz, Esmeraldas, and lower Mugrosa formations thin toward the west. Strata of the upper part of the Mugrosa Forma-
tion and the Colorado Formation thin toward the east instead. This eastward thinning correlates with late Oligocene to early middle Mio-
cene growth strata of the Provincia and Lisama anticlines (e.g., Figures 11 and 12) and suggests synsedimentary deformation of the eastern
flank of the Nuevo Mundo syncline (forelimb of the Los Cobardes anticline). See text for further discussion.

rotation of the associated folds (e.g., Anadón et of the upper boundary of the synorogenic strata of the Nuevo Mundo Syncline axial surface
al., 1986; Ford et al., 1997). of the Provincia anticline varies along its axis from near vertical upward to a moderate dip
Growth unconformities grade out away from from angularly unconformable in the zone of to the west (Fig. 10) resulted from folding of
the Provincia and Lisama folds into correla- highest structural relief (Fig. 12A), where ero- the originally nonparallel upper Mugrosa and
tive conformities. These compound surfaces, sion reached the lower Mugrosa Formation, Colorado beds (e.g., Ford et al., 1997) during
locally unconformable but elsewhere conform- to paraconformable toward the plunging tips younger Pliocene-Pleistocene deformation,
able, record the competition between vertical of this fold (Fig. 12B). The upper Mugrosa which led to the present configuration of the
accumulation of sediment and coeval structural and Colorado formations also thin toward the Nuevo Mundo syncline.
uplift, the rates of which varied through time. eastern flank of the Nuevo Mundo Syncline In contrast with Figure 7B, Figure 7C shows
Folding first caused displacement of sedimen- (i.e., forelimb of the Los Cobardes anticline; two younger Plio-Pleistocene faults. The west-
tation toward the basinal areas. Later, fluvial Fig. 10), which indicates that initial folding of ernmost fault is a low-angle thrust that ramps
deposits onlapped the growing folds as accu- the Los Cobardes was also contemporaneous across Cretaceous and Tertiary rocks, has its
mulation rates overcame uplift rates. The nature with the deposition of these units. The bending tip in the Mugrosa Formation, and merges with

Geological Society of America Bulletin, May/June 2005 559


GÓMEZ et al.

a steeper ramp that cuts across the basement


farther east. Two hanging wall anticlines are
associated with this fault. The Guayabito anti-
cline, to the west, is a leading fault-propagation
fold, whereas the Opón anticline, to the east,
is a fault-bend fold that formed in response to
the change in dip of the underlying fault ramp.
The La Salina fault cuts the Opón anticline and
has an approximate throw of 8 km. This reverse
fault is the youngest tectonic feature in the cross
section, as it is not folded by underlying fault
movements and truncates older folds (e.g., with
flanking Eocene-Oligocene growth strata) along
the length of the basin (Gómez et al., 2003).

SUMMARY OF THERMAL-HISTORY
CONSTRAINTS ON MMVB EVOLUTION

In this section, we summarize results of


thermochronological studies based on modeling
of detrital apatite-fission-track ages and track
length distributions, and apatite compositions
(Table 3), combined with vitrinite reflectance
analyses. These studies and techniques are fully
addressed elsewhere (Gómez, 2001; Green et
al., 1989). Apatite-fission-track and vitrinite
reflectance (Ro) samples were collected from
Mesozoic and Tertiary sandstones and mud-
stones in the Los Cobardes anticline (eastern
flank of the Nuevo Mundo syncline), La Cira-
Infantas oil field (Infantas-1613 well), and the
northernmost MMVB (Catalina-1 well; Figs. 2,
7B, and 7D).
Vitrinite reflectance indicates that mature
to overmature Cretaceous strata of the Los
Cobardes anticline forelimb (Ro = 0.6%–1.53%,
Ecopetrol-Esso, 1994; Fig. 7B) were heated
above the oil window and reached maximum
paleotemperatures between 110 and 170 °C
(e.g., Burnham and Sweeney, 1989). Modeling
of apatite-fission-track ages and length distribu-
tions suggests that two cooling events, ascribed
to uplift and erosion, affected this locality. The
apatite-fission-track parameters of all Creta-
ceous and Paleogene samples are consistent
with cooling from from peak paleotemperatures
in the order of 80–110 °C commencing some-
time during the past 5 m.y., as constrained by
sample RS-SS-2 (Table 3). With geothermal
gradients between 20–30 °C/km and a surface
temperature of 30 °C, this thermal episode may
reflect erosion of ~3 km of sedimentary section.
The Ro paleotemperatures of 110–170 °C were
most likely reached before an older cooling
episode, whose evidence is preserved by the Figure 11. Seismic line across the backlimb of the Lisama anticline. See Figure 2 for loca-
Cretaceous samples Tablazo-1-SS and Umir-SS tion. A wedge of growth strata characterized by divergent reflectors of the upper Mugrosa
(Fig. 7B, Table 3). These samples require cool- and Colorado formations fringes the backlimb of the Lisama anticline. These strata reveal
ing through ~120 °C sometime during the late that formation of this fold happened during the late Oligocene to early middle Miocene time.
Cretaceous-Paleogene time window 80–30 Ma Synorogenic geometries of the Lisama anticline were first found by geologist M. Suárez
owing to removal of 3–4 km of sedimentary (personal commun., 1999).

560 Geological Society of America Bulletin, May/June 2005


A B

Geological Society of America Bulletin, May/June 2005


SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

Figure 12. Seismic lines with examples of upper Oligocene to lower middle Miocene growth strata along the Provincia anticline. See Figure 2 for locations. The upper part of
the Mugrosa Formation and the Colorado Formation constitute a wedge-shaped unit of divergent reflectors that thicken away from the Provincia anticline axis. Packages of
preorogenic and postorogenic parallel reflectors are located above and below this synorogenic unit. Relationship between growth and post-growth strata changes from (A)
angular to (B) concordant toward the southern plunging tip of this fold. The lower surface of the upper Mugrosa-Colorado synorogenic wedge is a surface of onlap. Other

561
minor unconformities exist within this unit.
562
A B
TABLE 3. APATITE-FISSION-TRACK ANALYTICAL RESULTS FOR SAMPLES OF THE EASTERN FLANK OF THE NUEVO MUNDO SYNCLINE (i.e., LOS COBARDES ANTICLINE FORELIMB)
Sample number Stratigraphic unit Number ρs × 10 6 ρI × 105 ρd × 105 Fission track Chi2 Uranium Mean track Number Maximum Onset of Maximum Onset of
of grains (tracks/cm2) (tracks/cm2) (tracks/cm2) age (%) content length of tracks paleotemp cooling paleotemp cooling
(Ma ± 1 S.E.) (ppm) (µm ± 1 S.E.) measured (ºC) (Ma) (ºC) (Ma)
RS-SS30 Esmeraldas 9 1.596 2.962 1.128 118.2 ± 10.1 <1 30 11.18 ± 0.53 10 >110 45–0

(244) (453) (1733) 65.8 ± 30.1
RS-SS26 Esmeraldas 5 0.471 3.546 1.059 27.6 ± 5.3 8 38 10.24 ± 2.07 6 >110 45–0
(32) (241) (1707)
RS-SS21 Esmeraldas 4 0.633 1.966 1.066 67.1 ± 11.4 2 21 9.74 ± 74 15 (>100 45–15)† 80–100 15–0

(47) (146) (1707) 73.5 ± 20.2
RS-SS17 La Paz 15 0.401 2.103 1.074 40.1 ± 4.1 <1 22 11.62 ± 0.32 11 (>110 56–30)† 80–100 15–0

(125) (655) (1707) 40.3 ± 7.7
RS-SS14 La Paz 8 0.356 2.878 1.081 26.2 ± 4.2 97 30 8.94 ± 1.36 4 (>110 56–20)† 95–110 15–0
(45) (364) (1707)
RS-SS10 La Paz 26 0.689 3.208 1.089 45.8 ± 3.2 <1 34 11.56 ± 0.19 132 (>110 56–20)† 80–100 10–0
(318) (1480) (1707) 40.8 ± 5.1‡
RS-SS8 La Paz 4 0.130 1.112 1.096 25.1 ± 8.9 5 12 no confined tracks (>110 56–0)†
(9) (77) (1707)
RS-SS5 La Paz 28 0.386 1.822 1.104 45.8 ± 3.2 1 19 10.52 ± 0.24 118 (110–120 56–20)† 90–120 10–0

(317) (1496) (1707) 45.5 ± 4.6
GÓMEZ et al.

RS-SS2 Lisama 20 0.131 1.506 1.111 19.0 ± 2.3 <1 15 9.85 ± 0.46 21 (>110 65–30)† 100–110 5–0
(77) (883) (1707) 22.3 ± 4.2‡
RS-SS1 Lisama 25 0.308 1.646 1.119 40.9 ± 3.3 <1 17 10.25 ± 0.39 51 (>110 65–30)† 90–110 15–0

(216) (1156) (1707) 36.0 ± 5.3
Umir-SS Umir 20 0.350 1.738 1.472 56.1 ± 4.9 44 13 12.73 ± 0.40 23 >120 80–30 50–90 30–0

Geological Society of America Bulletin, May/June 2005


(Cretaceous) (169) (840) (2335)
Tablazo-1-SS Tablazo 21 1.140 3.560 1.488 90.0 ± 6.4 <1 27 11.14 ± 0.28 24 >110 124–15 80–95 15–0
(Cretaceous) (305) (952) (2335) 65.5 ± 17.1‡
Santos-SS Santos 13 1.534 1.959 1.504 220.2 ± 18.5 <1 15 12.08 ± 0.28 15 (<100 145–80)† 70–80 80–0
(Cretaceous) (278) (355) (2335) 234.1 ± 50.3‡
Giron-SS Giron 2 0.167 2.565 1.512 18.7 ± 7.9 63 19 no confined tracks (>110 178–0)† >110 30–0
(Jurassic) (6) (92) (2335)
Notes: ρs—spontaneous track density in internal surface. ρi—induced track density in mica external detector. ρd—track density in mica external detector. Brackets show number of tracks counted. S.E.—
standard error; Chi2—Chi-squared probability. Ages calculated using dosimeter glass CN5, with a zeta of 380.4 ± 5.7 for samples Umir-SS, Tablazo-1-SS, Santos-SS and Giron-SS. Ages calculated using
dosimeter glass CN5, with a zeta of 392.9 ± 7.4 for all other samples.

Episodes shown in parentheses are allowed, but not required by the apatite-fission-track data.

Central age, used where sample contains a significant spread of single grain ages (Chi2<5%).
SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

overburden. Apatite-fission-track parameters in in the Paleocene Lisama Formation (Fig. 13A; Bridge, 1995) and transported sediment toward
other samples allow this event but do not require Campbell, 1968). Growing marine influence the northeast to the Maracaibo basin (Forero,
it. As clarified in the next section, the oldest toward the north is revealed by marine fauna 1974). The Cantagallo sandstones accumulated
cooling event can be attributed to formation found in the northernmost MMVB (Lebrija-1 simultaneously west of the Infantas paleohigh
of the MMVU during the early Eocene. The well, Ramírez, 1988; Fig. 13A) and Paleocene (Suárez, 1997), within the paleo Central Cordil-
youngest episode of cooling correlates well shallow marine deposits and coal bearing layers lera catchment area.
with Eastern Cordillera Pliocene deformation, in the Ranchería Basin (Cáceres et al., 1980). As illustrated by the subcrop pattern of the
also documented by cross sections. The Ranchería Basin was the northeastern MMVU (Fig. 8A), a vast area of Jurassic and
According to vitrinite reflectance data, the continuation of the Paleocene MMVB prior to Cretaceous rocks was exhumed during the early
maturity of Cretaceous oil-source rocks seems left-lateral movement of the Bucaramanga fault Eocene propagation of the Central Cordillera
to decrease toward the west of the Eastern (Campbell, 1968). into the MMVB. Jurassic rocks sourced rhyolite
Cordillera foothills and toward the north along Kilometer-scale uplift of the Central Cordil- and red-bed clasts to the basal La Paz Formation
the MMVB. The Cretaceous rocks drilled by lera is the only explanation for the extensive conglomeratic beds. The area of exposed Creta-
the Infantas-1613 well (e.g., Fig. 2) are early to occurrence of deformed crystalline basement ceous rocks in the slope of the Central Cordil-
middle mature (Ro = 0.44%–0.8%) and reached and Mesozoic rocks underneath the MMVU lera decreased through time due to erosion and
maximum paleotemperatures on the order of (Figs. 1B, 7, and 8A). In addition to the reported sedimentary overlap, which is reflected in fewer
125 °C. Those of the northernmost MMVB are magnitudes of unroofing derived from zircon- sedimentary grains in the upper La Paz and
early mature (Ro = 0.52%–0.56%), as mea- fission-track data, remnants of Jurassic and Cre- Esmeraldas formations. Cretaceous and Jurassic
sured in the Catalina-1 well (Fig. 7D) and were taceous rocks in the Central Cordillera (Fig. 1B) grains may have been shed also by the Arcabuco
never heated above ~115 °C. The present depth and subsurface of the MMVB indicate sedimen- anticline and its continuation into the Santander
ranges of the Cretaceous sections of the southern tation and subsequent erosion of 5–10 km of Massif in restored position as suggested by
Infantas-1613 and northern Catalina-1 wells are Mesozoic sedimentary rocks before deposition regional paleogeography (Gómez, 2001). The
1303–3262 m and 695–1950 m, respectively. of onlapping middle Eocene-Neogene deposits. subtle unconformity between the La Paz-Can-
The variation in maximum paleotemperatures The magnitude of unroofing underneath the tagallo sandstones and Esmeraldas formations
between these two wells may be partially MMVU decreases toward the east (e.g., Figs. 7A (Fig. 14) records tilting of the terminal La Paz
explained by the history of onlap of the MMVU, and 7B) in the direction of decreasing paleoalti- depositional surface toward the west (west of
which shows delayed burial of the northernmost tude of the early Paleogene Central-Cordillera the Infantas paleohigh) and toward the east
MMVB by a thinner sedimentary pile (Fig. 7D). uplift. Thermal history constraints reveal that (east of Infantas paleohigh), which most prob-
3–4 km of sedimentary section were eroded ably resulted from uplift of the Infantas high.
DATA INTEGRATION AND from the Nuevo Mundo syncline area sometime Strata of the overlying Esmeraldas Formation
INTERPRETATION OF MMVB during the 80–30 Ma time window (Gómez, onlapped the composite La Paz/Esmeraldas-
CENOZOIC HISTORY 2001). At this locality, the MMVU represents MMVU unconformity and buried the Infantas
an early to middle Eocene sedimentary hiatus paleohigh (Fig. 14).
(~15 m.y. long; e.g., Fig. 3) and is the only major The Esmeraldas Formation in the western
Paleocene–Early Eocene unconformity to which such a magnitude of ero- MMVB is a stack of sandstones whose facies
sion can be ascribed. Therefore, we attribute this are similar to those of the Cantagallo sandstones
Several data sets along the MMVB indicate denudation to propagation of Central Cordillera (Suárez, 1997). Thus, the Esmeraldas Formation
that Central Cordillera kilometer-scale uplift uplift into the MMVB during the early to Middle may also represent a high-gradient piedmont
progressed from south to north during late Cre- Eocene (55 to ca. 45 Ma; e.g., Schamel, 1991). deposit in the western side of the MMVB that
taceous to early Eocene times (Campbell, 1968). changed eastward to the lower gradient flood-
Fission-track ages of zircons of the Central Cor- Middle Eocene–Early Oligocene plain and channel deposits studied in the Nuevo
dillera Mariquita Stock, west of the southern- Mundo Syncline (Fig. 13C). Paleoflow mea-
most MMVB (Fig. 1B), indicate cooling below The relief of the Central Cordillera has been surements indicate that the northern MMVB
250 ± 50 °C and erosion of 7–13 km of rock col- smoothed since middle Eocene as erosion rivers continued flowing toward the Maracaibo
umn since the Campanian (Gómez et al., 2003). replaced expanding deformation as the domi- Basin. Sedimentation of the Los Corros fos-
Cooling below 250 ± 50 °C in the northern nant local process; the proof of this process is sil horizon may represent low gradient of the
Central Cordillera happened essentially during the MMVU, which is a pediment surface that Esmeraldas river system as an adjustment to a
the Paleocene according to Toro’s (1999) fis- resulted from erosional retreat of cliffs and contemporaneous sea-level highstand (Fig. 14),
sion-track study on zircons from the crystalline taluses (Gómez et al., 2003). The boundary whose impact propagated from the Maracaibo
basement near Medellín. In the southernmost between the Central Cordillera and MMVB Basin at approximately the Eocene-Oligocene
MMVB (Guaduas syncline, Fig. 1B), a coarsen- began to move progressively westward and boundary (33–34 Ma; e.g., Haq et al., 1988).
ing upward sequence of Maastrichtian marine northward, as the middle Eocene to Neogene
shales to Maastrichtian braid-delta deposits to sedimentary fill onlapped the residual Central Oligocene–Early Middle Miocene Evolution
Paleocene alluvial-fan conglomerates records Cordillera pediment surface (the MMVU;
transformation of this region from a marine Figs. 13B and 13C). The upper La Paz amal- A boundary of greater stratigraphic and tec-
basin into a piedmont area (Gómez, 2001). gamated sandstones, which are restricted to tonic importance than the Los Corros, Mugrosa,
Northward propagation of Central Cordillera the eastern flank of the Nuevo Mundo syncline, and La Cira horizons exists within the Mugrosa
uplift is recorded in the northern MMVB by mark the position of the main fluvial channels, Formation, which correlates with initial defor-
delayed change to continental facies and Cen- which wandered within areas of maximum sub- mation of the Eastern Cordillera. In the Eastern
tral Cordillera provenance, both of which occur sidence of the alluvial plain (e.g., Mackey and Cordillera foothills, this boundary is a growth

Geological Society of America Bulletin, May/June 2005 563


564
N N
GÓMEZ et al.

Geological Society of America Bulletin, May/June 2005


PA: Provincia anticline
LA: Lisama anticline
LC: Los Cobardes anticline
AA: Arcabuco anticline

Figure 13. Synthesis of MMVB paleogeography through the Cenozoic. (A) Paleocene, (B) middle to late Eocene, and (C) late Eocene–earliest Oligocene. The contoured depth
to the MMVU (from Fig. 8A) is also shown in (B) and (C) and reflects the mountain drainage that brought sediment from the Central Cordillera to the MMVB depocenters.
(D) Late Oligocene to early middle Miocene and (E) late Miocene. These palinspastic maps restore ~110 km of strike-slip movement along the Bucaramanga fault and the
tectonic shortening across the Eastern Cordillera (e.g., Campbell, 1968; Forero, 1974; Dengo and Covey, 1993; Gómez, 2001). See text for sources of data and discussion.
SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

W E
Lisama-Provincia Nuevo Mundo

250 m
Cantagallo fault anticlines syncline

0m
0
Pliocene
L Unconformities at crest
10 of Lisama-Provincia
Oligoc. Miocene

M anticlines

Growth
strata
20 E
L

Long-term eustatic Sea Level Curve


30 E
Age (Ma)

L
40 La Paz-Esmeraldas
Eocene

M Cantagallo unconformity
sandstones
50
E MMVU
60 L Infantas paleohigh
Pal.

70 Maast.
Camp.
0 20 km GU: Growth unconformity
CC: Correlative conformity
Mollusk fossils Alluvial fan sandstones
Lacustrine deposits Overbank deposits
Delta and coastal plain deposits Fluvial-channel sandstones
Shallow marine mudstones Lacuna
MMVU: Middle Magdalena Valley unconformity
Figure 14. Chronostratigraphic diagram illustrating the temporal relations among unconformities and sedimentary facies of the northern
MMVB. The time gap associated with the MMVU decreases toward the east of the MMVB to a minimum of 10–15 m.y. in the Nuevo Mundo
syncline area. The gray rectangles superimposed on the long-term eustatic sea-level curve (right side of diagram, Haq et al., 1988) highlight
the correlations between sea-level highstands and occurrence of the Los Corros and La Cira fossil horizons. See text for discussion. Geologic
time scale after Berggren et al. (1995).

unconformity at the base of growth strata formed deformation with similar vergence in the Perijá uous hills of the Eastern Cordillera. Drainages
of the upper Mugrosa and Colorado formations, Range is revealed by structural studies (Kellogg, were no longer able to drain water and sedi-
which flank the Nuevo Mundo syncline and 1984) and is also constrained by late Oligocene- ment as effectively as before the emplacement
the Provincia and Lisama anticlines. The intra- early Miocene apatite-fission-track cooling ages of such a barrier, which caused aggradation of
Mugrosa surface becomes a correlative con- (Shagam et al., 1984). The La Salina fault (East- fluvial deposits in the MMVB; the combination
formity toward the central parts of the MMVB ern Cordillera) and Tigre fault (Perijá Range; of low regional gradient and local folding is
(Fig. 14). Initial deformation of the Eastern Cor- Kellogg, 1984) appear to belong to the same expressed by excellent preservation of small-
dillera foothills was rooted in incipient inversion structural trend in restored position. Widespread scale sedimentary features and fine-grained
of the La Salina normal fault and also appears late Oligocene to early Miocene deformation facies in the upper Mugrosa Formation (Fig. 5)
to correlate with a major change in paleoflow to the northeast of the MMVB may have been and accumulation of local lacustrine deposits
direction in the La Cira-Infantas area. In this linked to initial strike-slip activity of the Santa composed exclusively of fresh-water mollusks
area, early Oligocene drainage remained ori- Marta-Bucaramanga fault, which has since then (Mugrosa fossil horizon; e.g., Burbank et al.,
ented toward the north while the lower part of the been accommodated by reverse faulting and 1996; see also Olaya, 1997; Ramón, 1998). The
Mugrosa Formation accumulated, but changed inversion of Mesozoic grabens in the Santander early to early middle Miocene is also a time of
toward the southeast during sedimentation of the Massif and Eastern Cordillera (Kovas et al., global eustatic sea-level rise (Fig. 14; Haq et
upper Mugrosa Formation (Fig. 13D). 1982; Toro, 1991; Pindell et al., 1998). al., 1988), which may have contributed to the
The MMVB history fits well with broader We interpret that the Eastern Cordillera-Perijá tectonically enhanced accommodation of the
regional relations (Fig. 13D). In restored posi- Range-Santander Massif structural barrier MMVB. The adjustment of the MMVB gradi-
tion, the late Oligocene-early Miocene MMVB closed the connection between the MMVB and ent profile to the ensuing highstand is recorded
anticlines are part of a larger morpho-structural the Maracaibo Basin and caused a major diver- by the early middle Miocene deposits of the
unit whose northeastward continuation was sion in fluvial drainage toward the SE to the La Cira fossil horizon, which contain slightly
located in the Perijá Range. Simultaneous Llanos Basin, across valleys between discontin- brackish water mollusks.

Geological Society of America Bulletin, May/June 2005 565


GÓMEZ et al.

High percentages of feldspar in the sand- 1988; Gómez, 2001). We interpret this rotation as Eastern Cordillera is related to the mechanics
stones of the Mugrosa and Colorado formations the MMVB response to Plio-Pleistocene Eastern of thick-skinned deformation or to collisional or
reflect deeper levels of erosion and enlarged Cordillera tectonic loading and Central Cordillera Carribean tectonics. However, as noted below,
exposure of Mesozoic granodioritic plutons in erosional unloading. The tilting directions, south- similar transfers and timing exist widely in the
the Central Cordillera (Mesa, 1995). Cretaceous eastward in the northern MMVB and eastward Andean system, which suggests that the major
rocks of the Cáchira arch remained exposed in the southern MMVB, reflect the distribution attributes of the MMVB-Llanos basin system
until the Cáchira arch was buried by the upper- of Eastern Cordillera topography, whose width might be typical of the entire Andean foreland.
most 100 m of the Colorado Formation strata of and altitude are the largest between 5.5º and 7º N, As noted earlier, published interpretations
likely early middle Miocene age. Only at this and diminish to the north in the Santander Massif emphasize the role of late Cretaceous to Eocene
time was a connection established between the (e.g., Fig. 1A). accretion of the Western Cordillera (Barrero,
MMVB and the Lower Magdalena Valley Basin 1979; Etayo-Serna et al., 1983), which drove
and not 18 m.y. earlier during the early Oligo- DISCUSSION deformation and uplift of Colombia’s Central
cene as interpreted by Villamil (1999). Simul- Cordillera. In contrast, it is well documented
taneous marine sedimentation in the Lower In this section, we discuss the relevance of our that there has been no terrane accretion along
Magdalena Valley Basin has been documented results with regard to two major aspects: (1) vari- the Central Andean margin during the Creta-
by Duque-Caro (1980). ations from standard foreland basin behavior, ceous and Cenozoic (Jordan and Gardeweg,
and (2) comparisons with the Central Andes. 1989; Beck et al., 1994). Nevertheless, a strong
Late Middle Miocene-Late Miocene Two major related features of the MMVB are not tectonic similarity exists. In the Central Andes
Evolution widely discussed in treatments of typical retro- (Fig. 15), late Cretaceous to Paleocene collapse
arc foreland basin systems. These are the major of Mesozoic backarc basins triggered the first
The late middle to late Miocene is a time diachronous unconformity that subdivides the major stage of shortening-related Andean uplift,
of another major MMVB readjustment. Paleo- Cenozoic column, the MMVU, and the impact dominated by inversion tectonics (Mpodozis
drainage direction changed toward the north of thick-skinned deformation on long-term evo- and Ramos, 1990). One could argue that this
as the result of progressive confinement of lution of the basin system. Unconformities are collapse of the Chilean continental margin is
the MMVB between the Central and Eastern of course common in continental basins; within fundamentally analogous to the Western Cor-
Cordilleras (Hoorn et al., 1995) and north- foreland basin systems, wedgetop (piggyback) dillera accretion to the Colombian Andes. The
ward overlap of the Cáchira Arch (Fig. 13E). basins, the distal sectors of foredeep basins, and similarity may be masked by the difference in
Metamophic and igneous clasts most probably zones adjacent to or above a forebulge, which are relative modern locations caused by younger
were derived from the Central Cordillera; most especially prone to unconformity development Central Andean tectonic erosion, such that the
of the Central Cordillera Cretaceous sedimen- (Jordan, 1995; DeCelles and Giles, 1996). Both Colombian uplift (Central Cordillera) is cur-
tary cover had already been eroded or buried episodic tectonic activity and eustatic sea-level rently in the core of the Andean mountains,
under MMVB onlapping sediments. The large changes are common reasons for unconformity whereas the Chilean uplift (Domeyko Range)
percentages of sedimentary clasts of the Real development. Yet the large-scale features of lies now in the low elevation Neogene forearc.
Group were mostly shed by Eastern Cordillera the MMVU, captured by onlap across a major Similarities of the timing of Northern Andean
sources such as the crest of the Los Cobardes mountain front pediment, do not fit those molds. and Central Andean tectonic evolution outweigh
anticline, the Santander Massif, and central In particular, the MMVU traverses the proximal their differences. We have shown that there was
parts of the Eastern Cordillera, which were to distal sectors of the early Cenozoic basin; its kilometer-scale uplift of the Central Cordillera
being deformed contemporaneously (Shagam stratigraphic hiatus decreases toward the east but between late Cretaceous and early Eocene time
et al., 1984; Dengo and Covey, 1993). Strata of lasted at least 15 m.y. at any one location. The (see also Gómez et al., 2003). In this time inter-
the Real Group were deposited in the area of the MMVU is not tied to fluctuations in a generally val in what is now the Chilean forearc region, the
Eastern Cordillera foothills, although erosion steady shortening history but rather to the termi- Domeyko Range was inverted, forming the core
subsequently stripped them from areas south of nation of deformation in one thick-skinned belt of the Andean ranges with kilometer-scale uplift
the Nuevo Mundo syncline. Epiclastic deposits (the Central Cordillera) and transferal of defor- but only modest horizontal shortening (Fig. 15).
found in the upper Real Group were sourced by mation to a parallel but separate thick-skinned The middle Eocene and Oligocene record of the
andesitic volcanism in the Cauca Valley (e.g., zone (the inverted Mesozoic basin in the Eastern MMVB basin and neighboring ranges is espe-
Etayo-Serna et al., 1983; Thouret, 1989). Cordillera). The kilometer scale of erosional cially marked by the MMVU, with onlap across
downcutting across the MMVU is not appropri- the pediment cut above the Central Cordillera
Pliocene-Recent Evolution: Mesa Group ate to forebulge uplift (hundreds of meters; e.g., uplift. The Chilean late Paleocene to Eocene
and Regional Context Jordan, 1995). paleogeographic record is less well resolved.
The MMVB-Llanos basin system is unusual However, a major feature is the long-distance
The Pliocene Mesa Group sedimentation cor- among long-lived foreland basin systems in transport of large volumes of detritus eastward
relates with reactivation of the La Salina fault its connection to thick-skinned uplifts. The from Domeyko Range and accumulation across
system, which led to the present geometries of preconditioning of the continental lithosphere the eastern flank of that paleo-uplift across a
the Nuevo Mundo syncline and the Los Cobardes by a magmatic arc (Central Cordillera) and rift major erosional unconformity throughout its
anticline. Thermal history constraints, discussed system (Eastern Cordillera) may have reduced foreland basin, the Salar de Atacama basin.
before, reveal erosion of 3 km of sedimentary the opportunity for upper-crustal shortening to The northern Chile relations represent a situ-
section in the eastern flank of the Nuevo Mundo be accommodated by thin-skinned thrusting as ation potentially analogous to the development
syncline within the past 5 m.y. The MMVB is typical. Arguably, the reason for the major of the MMVU. In both the Central and Northern
Neogene strata dip eastward and are truncated eastward transfer in the locus of the shortening Andes, the Oligocene and early Miocene are
against the surface of the MMVB (Ramírez, and uplift from the Central Cordillera to the times of major reorganization of deformation

566 Geological Society of America Bulletin, May/June 2005


A

Geological Society of America Bulletin, May/June 2005


G
SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

Figure 15. (A) Regional location map for Salar de Atacama Basin in northern Chile; inset map shows location of regional map relative to South America. (B) to (G) Sequential
cartoon cross sections X–Y illustrating the evolution of the Salar de Atacama Basin since the deformation that occurred at approximately the Cretaceous-Paleocene bound-
ary. See Figure (A) for location of X–Y cross section. Strata in the same stratigraphic sequence as the age of the cross section are highlighted in gray. Evolution of the Salar de
Atacama Basin is potentially analogous to the MMVB. (B) Paleocene sequence H fills the tectonic subsidence generated by Cretaceous–earliest Paleocene folding and thrusting
(or reverse faulting) at the western margin of and within the Salar de Atacama Basin. (C) Eastward rotation and erosional truncation of sequence H. (D) Eocene sequence J
represents sediment shed from tectonically active highlands located tens of kilometers west of the Salar de Atacama Basin. (E) Oligocene–early Miocene sequence K fills an
extensional basin, whose western boundary was a set of east-facing normal faults. (F) Minor thrusting within the western sector of the Salar de Atacama Basin. The middle
to upper Miocene sequence L fills remnant accommodation space that was modified by localized subsidence. (G) Accommodation space for sequence M accumulation is con-
trolled by a complex combination of local deformation (e.g., SFS, Salar fault system) and basin-wide remnant accommodation. The evolution depicted by Figures (E) to (G) is
contemporaneous with segregation of the interior Altiplano basin from the exterior sub-Andean foreland basin, to the east of the Salar de Atacama basin. See text for further

567
discussion. Figure courtesy of Jordan, T.E., Mpodozi, C., Muñoz, N., Blanco, N., Pananont, P., and Gurdewes, M.
GÓMEZ et al.

International provided rock samples from the


and related basins. In the MMVB, this was the provoked by Eastern Cordillera loading. Proof Infantas-1613 and Catalina-1 wells, respectively,
time of shifting drainage directions (northeast- of Eastern Cordillera deformation is provided for AFTA and Ro studies. We also thank Edgar
ward, then eastward, and finally northward), in by growth strata, thermal history, and sediment Baquero, Mario Suárez, Juán Pablo Reyes, Iván
response to initiation of widespread shortening provenance. The subtle unconformity between Olaya, Juán C. Ramón, Timothy Cross, and Carlos
and gradual uplift in the Eastern Cordillera. the La Paz and Esmeraldas (Fig. 14) represents Jaramillo for helpful discussions. Critical reviews by
Peter De Celles, Paul Heller, Brian Horton, Timothy
Similarly, in the Central Andes, the major Oli- a short-term base-level fall overprint on regional Lawton, Wilfried Winkler, Yildirim Dilek, and Allen
gocene foreland basin, located in what is now accommodation due to minor reactivation of the Glazner helped us to improve this paper.
the Altiplano plateau, was followed in the late older Central Cordillera mountain front (i.e.,
Oligocene–early Miocene by major deforma- Infantas high). REFERENCES CITED
tion of its eastern flank, which led to segregation Initial deformation of the Eastern Cordillera
Allen, J.R.L., 1983, Studies in fluviatile sedimentation: Bars,
of the interior Altiplano basin from the exterior foothills along the MMVB was diachronous, bar complexes and sandstones sheets (Low-sinuosity
Subandean foreland basin (Horton et al., 2001; older in the south than in the north, as demon- braided streams) in the Brownstones (L.Devonian)
Elger, 2003). Lastly, the Neogene history of the strated by the ages of associated growth strata. Welsh Borders: Sedimentary Geology, v. 33, p. 237–
293, doi: 10.1016/0037-0738(83)90076-3.
Central Andes is characterized by long-term Middle Eocene-Oligocene synorogenic beds are Anadón, P., Cabrera, L., Ferrán, C., Marzo, M., and Riba, O.,
progression of thin-skinned crustal shorten- found in the Guaduas syncline of the southern- 1986, Syntectonic intraformational unconformities in
alluvial fan deposits, eastern Ebro Basin margins (NE
ing, coupled to steady evolution of a marginal most MMVB (e.g., Fig. 1B; Gómez et al., 2003). Spain), in Allen, P., and Homewood, P., eds., Foreland
Subandean foreland basin system as well as to Younger initial Eastern Cordillera deformation Basins: International Association of Sedimentologists
continued accumulation in relict foreland basins along the northern MMVB is shown by upper Special Publication 8, p. 259–271.
Aspden, J.A., McCourt, W.J., and Brook, M., 1987, Geo-
now trapped in the interior of the orogenic belt. Oligocene-early middle Miocene growth strata metrical control of subduction-related magmatism:
There is a noteworthy similarity between the (upper Mugrosa and Colorado formations) asso- The Mesozoic and Cenozoic plutonic history of
spatial and temporal pattern of Neogene uplift ciated with the Los Cobardes, Provincia, and Western Colombia: Journal of the Geological Society
of London, v. 144, p. 893–905.
of the Eastern Cordillera of Colombia, the for- Lisama anticlines (Fig. 14). Continued deforma- Barrero, D., 1979, Geology of the central Western Cordil-
mation of a marginal foreland basin (the Llanos tion in the Los Cobardes anticline during the late lera, west of Buga and Roldanillo, Colombia: Publi-
cación Especial, Ingeominas, v. 4, p. 75.
area), and persistence of an inherited interior Miocene is indicated by abundant Cretaceous Bayona, G.A., García, D.F., and Mora, G., 1994, La For-
basin (the MMVB) with the situation in the clasts in the Real Group. Intense Pliocene-Pleis- mación Saldaña: Producto de la actividad de estratovol-
Central Andes. Consequently, the major aspects tocene Eastern Cordillera deformation shown canes continentales en un dominio de retroarco, in Etayo
Serna, F., ed., Estudios Geológicos del Valle Superior
of the late Cretaceous through Cenozoic evolu- by structural relations and thermal history con- del Magdalena: Bogotá, Universidad Nacional de
tion of the Northern Andes seem fundamentally straints produced eastward flexural tilting of the Colombia-Ecopetrol Publicación Especial, p. I1–I21.
related to the evolution of the entire Andean MMVB. Shifting MMVB paleoflow, to the NE Beck, M.E., Jr., Burmester, R.E., Drake, R.E., and Riley, P.D.,
1994, A tale of two continents: Some tectonic contrasts
margin. We interpret that Caribbean tectonics during the middle Eocene–early Oligocene, to between the central Andes and North American Cordil-
and terrane accretions added extra complexity the SE during the late Oligocene–early Miocene, lera, as illustrated by their paleomagnetic signatures:
Tectonics, v. 13, p. 215–224, doi: 10.1029/93TC02398.
and modified the magnitudes of shortening and and back to the north since the late Miocene is Berggren, W.A., Kent, D.V., Swisher, C.C., II, and Aubry, M.,
uplift of Colombia’s Central and Eastern Cordil- consistent with evolution of the Eastern Cor- 1995, A revised Cenozoic geochronology and chro-
leras, but the overall result was modest variabil- dillera and consequent transformation of the nostratigraphy, in Berggren, W.A., Kent, D.V., Aubry,
M.P., Hardenbol, J., eds, Geochronology, time scales and
ity on the continental-scale Andean pattern. MMVB from a foreland into an interior basin. global stratigraphic correlation: Society for Sedimentary
Quaternary growth strata (Julivert, 1958, p. 20), Geology (SEPM) Special Publication 54, p. 129–212.
CONCLUSIONS stream offsets (Campbell, 1968), and shallow Bogotá, J., and Aluja, J., 1981, Geología de la Serranía de
San Lucas: Geología Norandina, v. 4, p. 49–55.
seismicity (Taboada et al., 2000) indicate con- Burbank, D.W., Meigs, A., and Brozovic, N., 1996, Interac-
The MMVB data document late Cretaceous- tinued Eastern Cordillera tectonic activity. Thus, tions of growing folds and coeval depositional systems:
Basin Research, v. 8, p. 199–223, doi: 10.1046/J.1365-
early Eocene uplift of the Central Cordillera the MMVB is likely still experiencing flexural 2117.1996.00181.X.
and subsequent transferal of deformation to adjustment contemporaneously with continued Burnham, A.K., and Sweeney, J.J., 1989, A chemical kinetic
the Eastern Cordillera, in a manner consistent formation of the MMVU in the present slopes of model of vitrinite reflectance maturation: Geochimicha
and Cosmochimica Acta, v. 53, p. 2649–2657, doi:
with Andean-margin thick-skinned evolution. the Central Cordillera. 10.1016/0016-7037(89)90136-1.
The MMVU is a regional pediment surface Cáceres, H., Camacho, R., and Reyes, J., 1980, Guide-
that records the Central Cordillera uplift and ACKNOWLEDGMENTS book to the geology of the Ranchería Basin: Bogotá,
Sociedad Colombiana de Geólogos y Geofísicos del
erosional histories. Accommodation space was This study was supported by grants and fellow- Petróleo, 19th field conference, April 18–19, 71 p.
large during the depositional phases separated ships from the Petroleum Research Fund (ACS-PRF Campbell, C.J., 1968, The Santa Marta wrench fault of
#32818-AC8), National Science Foundation (Faculty Colombia and its regional setting, in Fourth Caribbean
by this surface, as suggested by abundant Ceno- Geological Conference, 1965: Port of Spain, Trinidad,
zoic flood-plain lithofacies (e.g., Paola, 2000; Award to Women in Science and Engineering award
p. 247–261.
GER-9022811 to T.E. Jordan), and the Instituto Cediel, F., 1968, El Grupo Girón, una molasa mesozoica de
Fig. 14). The MMVB accommodation was most Colombiano para el desarrollo de la ciencia y la la Cordillera Oriental: Boletín Geológico del Ingeomi-
likely related to the evolution of the bounding tecnología (Colciencias). Cornell University, the nas, v. 16, no. 1–3, p. 5–96.
ranges because crustal loads produced by tec- Geological Society of America, the American Asso- DeCelles, P.G., and Giles, K.N., 1996, Foreland basin sys-
tonic activity cause subsidence (Jordan, 1995). ciation of Petroleum Geologists, Shell Oil Company tems: Basin Research, v. 8, p. 105–123, doi: 10.1046/
Foundation, Shell E&P Solutions, and Ecopetrol J.1365-2117.1996.01491.X.
Thus, accumulation of Paleocene strata can Dengo, C.A., and Covey, M.C., 1993, Structure of the Eastern
also contributed funding to this research. Ecopetrol,
be explained by tectonic subsidence and large Esso Colombiana, and Maxus Energy provided well Cordillera of Colombia: Implications for Trap Styles
and Regional Tectonics: American Association of Petro-
sediment supply related to uplift of the Central and seismic data. The geochronological and thermo- leum Geologists Bulletin, v. 77, p. 1315–1337.
Cordillera. In contrast, regional westward onlap chronological studies were conducted by Geotrack Dickey, P., 1992, La Cira-Infantas Field, Middle Magdalena
and eastward thickening of middle Eocene- International (AFTA studies), New Mexico Tech Basin, in Beaumont, E.A., and Foster, N.H., eds.,
(fission-track ages of volcanic interbeds), and Dr. Structural Traps VII: Tulsa, American Association of
Neogene strata during massive Central Cordil- Alan Cook (Keiraville Konsultants, Ro analyses). Petroleum Geologists Treatise of Petroleum Geology,
lera erosion are consistent with subsidence The Instituto Colombiano del Petróleo and Harken Atlas for oil and gas fields, p. 323–347.

568 Geological Society of America Bulletin, May/June 2005


SYNTECTONIC SEDIMENTATION IN THE MIDDLE MAGDALENA VALLEY BASIN

Duque-Caro, H., 1980, Geotectónica y evolución de la Jordan, T.E., 1995, Retroarc Foreland and related basins, Ramírez, R.E., 1988, Stratigraphy of the Tertiary of the Mid-
región noroccidental Colombiana: Boletín Geológico in Busby, C.J., and Ingersoll, R.V., eds., Tectonics of dle Magdalena Basin (Colombia), central and northern
Ingeominas, v. 23, p. 4–37. sedimentary basins: Malden, Massachusetts, Blackwell parts, [M.S. Thesis]: Austin, University of Texas, 199 p.
Duque-Caro, H., 1990, The Chocó Block in the northwest- Science, p. 331–362. Ramón, J.C., 1998, Sequence stratigraphic framework of
ern corner of South America: Structural, tectonostrati- Jordan, T.E., and Gardeweg, M., 1989, Tectonic Evolution of Tertiary strata and oil geochemical evaluation, Middle
graphic, and paleogeographic implications: Journal of the late Cenozoic Central Andes, in Ben-Avraham, Z., Magdalena Basin, Colombia [Ph.D. thesis]: Golden,
South American Earth Sciences, v. 3, p. 71–84, doi: ed., The Evolution of the Pacific Ocean Margin: New Colorado, Colorado School of Mines, 270 p.
10.1016/0895-9811(90)90019-W. York, Oxford University Press, p. 193–207. Restrepo, P.A., 1995, Late Precambrian to early Mesozoic
Ecopetrol-Esso, 1994, Geochemical database, Integrated Julivert, M., 1958, La morfoestructura de la zona de mesas al SW tectonic evolution of the Colombian Andes, based on
technical evaluation Santander Sector: Colombia, de Bucaramanga (Colombia S.A.): Boletín de Geología new geochronological, geochemical and isotopic data
Annex V, unpublished internal report, 168 p. Universidad Industrial de Santander, no. 1, p. 9–43. [Ph.D. thesis]: Tucson, University of Arizona, 189 p.
Elger, K., 2003, Analysis of deformation and tectonic his- Julivert, M., 1963, Los rasgos tectónicos de la región de Retallack, G.J., 1988, Field recognition of paleosols, in Rein-
tory of the Southern Altiplano Pateau (Bolivia) and la Sabana de Bogotá y los mecanismos de formación hardt, J., and Sigleo, W.R, Paleosols and weathering
their impotance for plateau formation [Ph.D. thesis]: de las estructuras: Boletín de Geología Universidad through geologic time: Principles and applications: Geo-
GeoForschungsZentrum Potsdam, Scientific Technical Industrial de Santander, no. 13–14, p. 5–102. logical Society of America Special Paper 216, p. 1–20.
Report, v. STR03, no. 05, p. 152. Kellogg, J.N., 1984, Cenozoic tectonic history of the Sierra Rueda, M., 1996, Palinoestratigrafía del Terciario de la
Etayo-Serna, F., Barrero, D., Lozano, H., Espinosa, A., de Perijá, Venezuela-Colombia, and adjacent basins, in Cuenca del Valle Medio del Magdalena: Instituto
González, H., Orrego, A., Ballesteros, I., Forero, H., Bonini, W.E., Hargraves, R.B., and Shagam, R., eds., Colombiano del Petróleo-Ecopetrol, technical report,
Ramírez, C., Zambrano, F., Duque, H., Vargas, R., The Caribbean-South America Plate Boundary and unpublished, 105 p.
Núñez, A., Alvarez, J., Ropaín, C., Cardozo, E., Gal- Regional Tectonics: Geological Society of America Schamel, S., 1991, Middle and Upper Magdalena Basins,
vis, N., Sarmiento, L., Albers, J., Case, J., Singer, D., Memoir 162, p. 239–261. Colombia, in Biddle, K.T., ed., Active Margin Basins:
Bowen, R., Berger, B., Cox, D., Hodges, C., 1983, Mapa Kovas, E.J., Rodgers, D.A., and Binger, S.H., 1982, Seismic American Association of Petroleum Geologists Mem-
de Terrenos Geológicos de Colombia: Publicaciones interpretation of back thrusts and a displacement tran- oir 52, p. 283–301.
Geológicas Especiales del Ingeominas 14, 235 p. fer zone between two en echelon thrust faults, Middle Servicio Geológico Nacional, 1966, Geología del Cuadrán-
Fabre, A., and Delaloye, M., 1982, Intrusiones básicas Cre- Magdalena Valley, in Quinto Congreso Latinoameri- gulo I-11, Cimitarra: scale 1:200,000, 1 sheet.
tácicas de la Cordillera Oriental: Geología Norandina, cano de Geología: Argentina, Actas, v. I, p. 565–582. Servicio Geológico Nacional, 1967, Geología del Cuadrán-
no. 6, p. 19–28. Macellari, C., 1988, Cretaceous paleogeography and depo- gulo H-11, Barrancabermeja: scale 1:200,000, 1 sheet.
Folk, R.L., 1974, Petrology of sedimentary rocks: Austin, sitional cycles of western South America: Journal of Shagam, R., Kohn, B.P., Banks, P.O., Dasch, L.E., Vargas,
Hermphill’s, 170 p. South American Earth Sciences, v. 1, p. 373–418. R., Rodríguez, G.I., and Pimentel, N., 1984, Tectonic
Ford, M., Williams, E.A., Artoni, A., Verges, J., and Mackey, S.D., and Bridge, J.S., 1995, Three-dimensional implications of Cretaceous–Pliocene fission track
Hardy, S., 1997, Progressive evolution of a fault- model of alluvial stratigraphy: Theory and application: ages from rocks of the circum-Maracaibo Basin
related fold pair from growth strata geometries, Sant Journal of Sedimentary Research, v. B65, p. 7–31. region of western Venezuela and eastern Colombia,
Llorenç de Morunys, SE Pyrenees: Journal of Struc- Mesa, A.G., 1995, Diagénesis y calidad del reservorio Campo in Bonini, W.E., Hargraves, R.B., and Shagam, R.,
tural Geology, v. 19, p. 413–441, doi: 10.1016/S0191- La Cira, Formaciones Mugrosa y Colorado, Valle Medio eds., The Caribbean-South American plate boundary
8141(96)00116-2. del Magdalena: Ecopetrol-Instituto Colombiano del and regional tectonics: Geological Society of America
Forero, O., 1974, The Eocene of northwestern South Amer- Petróleo, Informe Final, 44 p. Memoir 162, p. 385–412.
ica [M.S. thesis]: Tulsa, University of Tulsa, 81 p. Miall, A., 1988, Reservoir heterogeneities in fluvial sandstones: Suárez, M.A., 1997, Facies analysis of the Upper Eocene
Geotec Ltda, 1988, Geologic Map of Colombia, second edi- Lesson from Outcrop Studies: American Association of La Paz Formation, and regional evaluation of the Post-
tion: Geotec Ltda, scale 1:1,200,000, 1 sheet. Petroleum Geologists Bulletin, v. 72, p. 682–696. middle Eocene stratigraphy, Northern Middle Magda-
Geotec-Maxus, 1993, Stratigraphic and tectonic study of Miall, A., 1992, Alluvial deposits, in Walker, R., and James, lena Valley Basin, Colombia [M.S. thesis]: Boulder,
the Upper and Middle Magdalena Basins, Colombia: N., Facies Models, Response to Sea Level Change: Saint University of Colorado, 88 p.
preliminary report, unpaginated, unpublished internal Johns, Geological Association of Canada, p. 119–142. Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A.,
report, 31 p. Morales, L.G., Podesta, D.J., Hatfield, W.C., Tanner, H., Hervé, P., Harmen, B., Olaya, J., and Rivera, C., 2000,
Gómez, E., 2001, Tectonic controls on the late Cretaceous Jones, S.H., Barker, M.H., O’Donoghue, D.J., Mohler, Geodynamics of the northern Andes: Subductions and
to Cenozoic sedimentary fill of the Middle Magdalena C.E., Dubois, E.P., Jacobs, C., and Goss, C.R., 1958, intracontinental deformation (Colombia): Tectonics,
Valley Basin, Eastern Cordillera and Llanos Basin, General geology and oil occurrences of the Middle v. 19, p. 787–813, doi: 10.1029/2000TC900004.
Colombia [Ph.D. thesis]: Ithaca, New York, Cornell Magdalena Valley, Colombia: Tulsa, Habitat of Oil Thouret, J.C., 1989, Geomorfología y crono-estratigrafía
University, 619 p. Symposium, American Association of Petroleum del macizo volcánico Ruiz-Tolima (Cordillera Central
Gómez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., Geologists, p. 641–695. Colombiana), in Van Der Hammen, T., Díaz-Piedra-
Kelley, S., and Heizler, M., 2003, Controls on architec- Mpodozis, C., and Ramos, V.A., 1990, The Andes of Chile hita, S., and Alvarez, V.J., eds., Estudios de ecosiste-
ture of the late Cretaceous to Cenozoic Southern Mid- and Argentina, in Ericksen, G.E., Pinochet, M.T.C., mas tropandinos, La Cordillera Central Colombiana,
dle Magdalena Valley Basin, Colombia: Geological Reinemund, J.A., eds., Geology of the Andes and Transecto Parque Los Nevados, v. 3, segunda parte:
Society of America Bulletin, v. 115, no. 2, p. 131–147. its Relation to Hydrocarbon and Mineral Resources: Berlin, J. Cramer, p. 257–277.
Green, P.F., Duddy, I.R., Gleadow, A.J.W., and Lovering, J.F., Houston, Circum-Pacific Council on Energy and Min- Toro, J., 1991, The termination of the Bucaramanga fault in the
1989, Apatite Fission Track Analysis as a palaeotem- eral Resources, Earth Science Series, p. 59–90. Cordillera Oriental, Colombia, [abs.]: Vth International
perature indicator for hydrocarbon exploration, in Nae- Nuttal, C.P., 1990, A review of the Tertiary non-marine mol- Circumpacific terrane conference, Serie Comunicaciones,
ser, N.D., and McCulloh, T., eds., Thermal history of luscan faunas of the Pebasian and other inland basins Departamento de Geología, Facultad de Ciencias Físicas
sedimentary basins—methods and case histories: New of north-western South America: Bulletin of the British y Matemáticas, Universidad de Chile, no. 42, p. 226.
York, Springer-Verlag, p. 181–195. Museum of Natural History, v. 45, p. 165–371. Toro, G., 1999, Chronology of the volcanic activity and
Gutiérrez, M., 2001, Seismic characterization of reservoir Olaya, I.D., 1997, Seismic stratigraphic characterization of regional thermal events: A contribution from the
heterogeneities to improve recovery efficiency [Ph.D. the Lower Tertiary in the Cachira Paleohigh, Middle tephrochronology in the north of the Central Cordillera
thesis]: Stanford, California, Stanford University, 170 p. Magdalena Basin, Colombia [M.S. thesis]: Golden, Colombia, in 4th International Symposium on Andean
Haq, B.U., Hardenbol, J., and Vail, P.R., 1988, Mesozoic Colorado School of Mines. Geodynamics: Göttingen, Germany, IRD-ORSTOM,
and Cenozoic chronostratigraphy and cycles of sea- Paola, C., 2000, Quantitative models of sedimentary basin p. 761–763.
level change, in Wilgus, C.K., et al., eds., Sea-level filling: Sedimentology, v. 47, Suppl. 1, p. 121–178, doi: Van der Hammen, T., and Garcia, C., 1966, The Paleocene
changes: An integrated approach: Society of Economic 10.1046/J.1365-3091.2000.00006.X. pollen flora of Colombia: Leidse Geologische Med-
Paleontologists and Mineralogists Special Publica- Pilsbry, H.A., and Olsson, A.A., 1935, Tertiary fresh-water edelingen, v. 35, p. 105–114.
tion 42, p. 71–108. mollusks of the Magdalena embayment, Colombia, with Villamil, T., 1999, Campanian–Miocene tectonostratigra-
Hoorn, C., Guerrero, J., Sarmiento, G.A., and Lorente, M.A., Tertiary Stratigraphy by O.C. Wheeler: Proceedings of phy, depocenter evolution and basin development of
1995, Andean tectonics as a cause for changing drainage pat- the Academy of Natural Sciences of Philadelfia, v. 87, Colombia and western Venezuela: Palaeogeography,
terns in Miocene northern South America: Geology, v. 23, p. 7–39. Palaeoclimatology, Palaeoecology, v. 153, p. 239–275,
p. 237–240, doi: 10.1130/0091-7613(1995)0232.3.CO;2. Pindell, J.L., Higgs, R., and Dewey, J.F., 1998, Cenozoic doi: 10.1016/S0031-0182(99)00075-9.
Hopping, C.A., 1967, Palynology and the oil industry, in palinspastic reconstruction, paleogeographic evolu- Ward, D.E., Goldsmith, R., Jimeno, A., Cruz, J., Restrepo, H.,
Cenophytic palynology: Reviews of Palaeobotony and tion, and hydrocarbon setting of the northern margin of and Gómez, E., 1977, Mapa geológico del cuadrángulo
Palynology, v. 2, p. 23–48. South America in Pindell, J.L., and Drake, C.L., eds., H-12, Bucaramanga, Colombia: U.S. Geological Sur-
Horton, B.K., Hampton, B.A., and Waanders, G.L., 2001, Paleogeographic evolution and non-glacial eustasy, vey and Ingeominas, scale 1:100:000, 1 sheet.
Paleogene synorogenic sedimentation in the Altiplano northern South America: Society for Sedimentary
Plateau and implications for initial mountain building Geology (SEPM) Special Publication 58, p. 45–85. MANUSCRIPT RECEIVED BY THE SOCIETY 21 JULY 2003
in the Central Andes: Geological Society of America Rabe, E.H., 1977, Zur Stratigraphie des ostandinen Raumes REVISED MANUSCRIPT RECEIVED 18 FEBRUARY 2004
MANUSCRIPT ACCEPTED 27 MAY 2004
Bulletin, v. 113, p. 1387–1400, doi: 10.1130/0016- von Kolumbien: Giessener Geologische Schriften,
7606(2001)1132.0.CO;2. Lenz-Verlag, Giessen, no. 11, 223 p. Printed in the USA

Geological Society of America Bulletin, May/June 2005 569

Potrebbero piacerti anche