Sunteți pe pagina 1din 22

Downloaded from gsabulletin.gsapubs.

org on December 15, 2012

Geological Society of America Bulletin


Development of the Colombian foreland-basin system as a consequence of
diachronous exhumation of the northern Andes
Elas Gmez, Teresa E. Jordan, Richard W. Allmendinger and Nestor Cardozo
Geological Society of America Bulletin 2005;117, no. 9-10;1272-1292
doi: 10.1130/B25456.1

Email alerting services

click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when new


articles cite this article

Subscribe

click www.gsapubs.org/subscriptions/ to subscribe to Geological Society of


America Bulletin

Permission request

click http://www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA

Copyright not claimed on content prepared wholly by U.S. government employees within scope of
their employment. Individual scientists are hereby granted permission, without fees or further
requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent
works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms
to further education and science. This file may not be posted to any Web site, but authors may post
the abstracts only of their articles on their own or their organization's Web site providing the posting
includes a reference to the article's full citation. GSA provides this and other forums for the
presentation of diverse opinions and positions by scientists worldwide, regardless of their race,
citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect
official positions of the Society.

Notes

Geological Society of America

Downloaded from gsabulletin.gsapubs.org on December 15, 2012

Development of the Colombian foreland-basin system as a consequence of


diachronous exhumation of the northern Andes
Elas Gmez
Teresa E. Jordan
Richard W. Allmendinger
Nestor Cardozo
Department of Earth and Atmospheric Sciences, Cornell University, Ithaca, New York 14853, USA

ABSTRACT
This study addresses multiple controls
on foreland-basin accommodation and contributes to enhanced understanding of the
evolution of the northern Andes. The Middle
Magdalena Valley Basin (MMVB), Eastern
Cordillera, and Llanos Basin are part of a
Late CretaceousCenozoic foreland-basin
system, east of the Colombian Central Cordillera. Mechanical modeling indicates that the
primary control on complex distributions of
sedimentary thicknesses, facies, and unconformities was lithospheric flexure in response
to crustal loads from the Central and Eastern
Cordilleras. Shorter-wavelength folding and
paleoaltitude determined the local character
of strata. Our mechanical modeling consists
of the application of orogenic and sedimentary loads extracted from geologic data on
a continuous elastic lithosphere. The results
validate two major basin configurations. The
first configuration was a Maastrichtianearly
Eocene foreland basin coupled with Central
Cordillera uplift. Growth strata record continuous sedimentation in the Eastern Cordillera, whereas regional unconformities in the
Llanos Basin (distal foreland basin) reflect
isostatic adjustments of the basins amplitude
and wavelength to Central Cordillera episodic
uplift and tectonic quiescence. The second
major basin configuration was characterized
by Central Cordillera erosion since middle
Eocene times recorded by a regional pediment
surface. In the absence of Central Cordillera effective loading, loads from onlapping
sediments and Eastern Cordillera piggyback
sub-basins provoked postmiddle Eocene

Present address: Shell International Exploration


and Production Inc., E&P Solutions, 200 North
Dairy Ashford, Houston, Texas 77079, USA; e-mail:
elias.gomez@shell.com.

accommodation in the MMVB and Llanos


Basin. Intensified Eastern Cordillera uplift
during the Neogene produced basinal tilting
recorded by unconformities in the MMVB.
This study highlights the importance of assessing the causes of tectonic accommodation as a
foundation for interpretation of the evolution
of large foreland and intermontane basins.
Keywords: basin analysis, subsidence, unconformity, paleogeographic controls, Colombia,
Northern Andes.
INTRODUCTION
The Middle Magdalena Valley Basin
(MMVB), the now-uplifted Eastern Cordillera area, and the Llanos Basin belong to a
large Andean foreland-basin system east of the
Central Cordillera (Fig. 1), which formed in
response to Late CretaceousCenozoic convergent-margin tectonics. This region provides an
opportunity to study various scales of tectonic
controls on accommodation of sedimentary
basins coupled to large orogens. In this context,
the main objective of this paper is to explain the
mechanical causes of the complex distributions
of Maastrichtian-Cenozoic facies and unconformities that resulted from diachronous exhumation of the Central and Eastern Cordilleras. We
integrate numerous data sets and use mechanical
models to test the validity of interpretations of
tectonic accommodation. This study contributes
two main sets of results: first, knowledge of the
controls on basinal wavelengths and amplitudes,
which can be applied to foreland and intermontane basins elsewhere; and second, enhanced
understanding of the evolution of the northern
Andes region, as the chronologies of interaction
between tectonic subsidence, sedimentation, and
mountain deformation proposed in this paper are
substantially different from those envisioned by
previous works.

The attributes of the Colombian forelandbasin sedimentary fill, locally reaching 10 km


in thickness, can be grouped into two main scale
categories. The first category involves features
and stratigraphic changes that occur over horizontal scales of hundreds of kilometers, which
reflect regional isostatic responses to tectonic
and sedimentary loading and erosion. This
group includes an eastward change from continental sedimentation in the MMVB to coastal
environments that predominated over a large
portion of the Llanos Basin history and regional
unconformities in the MMVB and Llanos Basin
region. The second scale of attributes involves
features with extents of kilometers to tens of
kilometers such as growth unconformities and
local distributions of facies, associated with
shorter-wavelength synsedimentary folding.
This study is the continuation of recent
MMVB papers (Gmez et al., 2003, 2005),
summarized in a later section, which provide
data necessary to reconstruct the links between
sedimentary fill and mountain evolution.
The MMVB contains the best record of dual
development of the Central and Eastern Cordilleras. The strata in this basin record Maastrichtianearly Eocene eastward propagation of the
Central Cordillera mountain front, followed by
a record of middle EoceneNeogene erosion of
the Central Cordillera and simultaneous Eastern
Cordillera deformation. In order to investigate the effects of tectonic loading and stratal
accumulation on the final accommodation
histories from the MMVB to the Llanos Basin
across the Eastern Cordillera, we performed
the following basin-analyses steps, whose
results are described sequentially through this
paper: (1) quantitative assessment of subsidence derived from backstripped sedimentary
columns; (2) analyses of accommodation patterns, as retrieved from sedimentary thicknesses
and facies, and interpretations of mechanisms
of subsidence and genesis of unconformities;

GSA Bulletin; September/October 2005; v. 117; no. 9/10; p. 12721292; doi: 10.1130/B25456.1; 17 figures; 1 table; Data Repository item 2005148.

1272

For permission to copy, contact editing@geosociety.org


2005 Geological Society of America

Geological Society of America Bulletin, September/October 2005

MMVB

300 km

50

100 km

SE
Llanos Basin

South
America

Neogene sedimentary rocks


Jurassic-Paleogene sedimentary rocks
Pre-Mesozoic basement and
Mesozoic plutons

Peru

Eastern Cordillera

Ecuador

Colombia

Venezuela

Romeral
fault system

GS

Medell n

Bogot
FS BA
US

VA

NM
LC
LA

PA

AA

Thrust fault
Strike-slip fault
Anticline
Syncline

Metropolitan region

Southern MMVB
(Figure 2A)

LMVB, MMVB, UMVB:


Lower, Middle, and Upper
Magdalena Valley Basins

Pliocene-Quaternary deposits
Plio-Pleistocene volcanoes
Cenozoic sedimentary rocks
Cretaceous marine rocks
Jurassic red beds and volcanics
Pre-Mesozoic basement and
Mesozoic plutons

Boyac

Eastern Cordillera cross section (Figure 1C)

Chronostratigraphic sections
(Data Repository items)

Eastern Cordillera

LMVB

200 km

Central
Cordillera

100

Santa Marta
Massif

Figure 1. (A) Shaded relief map of Colombia and nearby regions in northwestern South America illustrating the general tectonic and physiographic setting of the Central
Cordillera, Middle Magdalena Valley Basin (MMVB), Eastern Cordillera, and Llanos Basin, which are the subject areas of this paper. (B) General geologic map of the MMVB,
adjacent Central and Eastern Cordilleras, and Llanos Basin with localities mentioned throughout this paper. Main map source: Geotec Ltda (1988). (C) Structural section
across the Eastern Cordillera simplified after Dengo and Covey (1993, see location in Fig. 1B).

NW

Panam

MM
VB

BF

UM
VB

Ra
n
Ba cher
sin a

ng
e
Pe
r
i
j R
a

nder
Santa sif
Mas

Ll
B ano
a
sin s

BF:Bucaramanga fault
PA: Provincia anticline
LA: Lisama anticline
NM: Nuevo Mundo syncline
LC: Los Cobardes anticline
AA: Arcabuco anticline
GS: Guaduas syncline
VA: Villeta anticlinorium
FS: Fusagasug syncline
US: Usme syncline
BA: Bogot anticline

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

1273

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

(3) testing of hypotheses by means of twodimensional mechanical modeling of subsidence; and (4) extrapolation of two-dimensional
modeling results to investigate broader regional
evolution of the northern Andes.
GEOLOGIC SETTING
The basins considered in this paper lie on
Proterozoic to Paleozoic continental basement
of South America, which is bounded to the west
by the Romeral fault system along the western
flank of the Central Cordillera (Etayo-Serna et
al., 1983, Fig. 1A herein). Mesozoic and Cenozoic calc-alkaline plutons intrude older metamorphic complexes of the Eastern and Central
Cordilleras as well as accreted oceanic terranes
to the west of the Romeral fault (Etayo-Serna et
al., 1983). Lithospheric stretching characterized
the MMVB and Eastern Cordillera areas during
the Mesozoic (Etayo-Serna et al., 1983). Triassic to Jurassic synrift red beds are exposed at the
core of Eastern Cordillera anticlines (Fig. 1B).
Three main rifted sub-basins persisted during
the Cretaceous to the east of the Central Cordillera. The Magdalena-Tablazo and the Cocuy
sub-basins (Fabre, 1983a, 1983b) were located
to the west and east of the present Santander
Massif and merged toward the south into the
Cundinamarca sub-basin at the present location
of Bogot and the Villeta anticlinorium (Fig. 1B
herein, Sarmiento, 1989). A marine transgression during the Cretaceous deposited a transgressive-regressive megasequence of mainly
shales and limestones (Macellari, 1988).
The Late CretaceousTertiary exhumation of
the Central and Eastern Cordilleras was linked
to the evolution of the western active margin
of South America. Late CretaceousEocene
oblique accretion of the Western Cordillera
caused northward propagation of uplift of the
Central Cordillera (Campbell, 1968; EtayoSerna et al., 1983). Compressional deformation
and tectonic inversion of Mesozoic grabens
in the Eastern Cordillera area were also initiated at that time and continued throughout the
Cenozoic (Julivert, 1963; Gmez, 2001). The
most intense pulse of Eastern Cordillera uplift
started at 12.9 Ma and has been attributed to
accretion of the Panam-Baud arc (Dengo and
Covey, 1993).
No estimates of orogenic shortening exist for
the Central Cordillera. This range may represent
a crustal-scale, positive flower structure (D.
Barrero, 1989, 1999, personal commun.). A
significant component of compressional deformation generated important uplift. The Central
Cordillera basement reaches altitudes of 3500 m
in places where it is not overlain by Pliocene
volcanoes. As explained later, the age of major

1274

deformation of the Central Cordillera at the


latitude of the MMVB is premiddle Eocene, as
evidenced by a major unconformity, the Middle
Magdalena Valley unconformity (MMVU).
The Eastern Cordillera increases its width
from south to north (1N to 7N, Fig. 1A).
North of 7N, the Eastern Cordillera structural
trend changes to NNW, and it also splits into
a NW branch, the Santander Massif and Perij
Range, and a NE branch, the Mrida Andes. The
Eastern Cordillera south of 7N is characterized
by oppositely verging fold and thrust belts,
which overthrust the MMVB and Llanos Basin
(Figs. 1B and 1C). Discrepant values of shortening between 68 and 170 km have been obtained
from regional cross sections at the same locality
and reflect different interpretations of the relative importance assigned to basement-involved
faulting versus thin-skinned styles of deformation (Figs. 1B and 1C; Colletta et al., 1990;
Dengo and Covey, 1993; Cooper et al., 1995;
Roeder and Chamberlain, 1995; Taboada et al.,
2000). The orientation of these cross sections
(120SE) is perpendicular to the structural trend
of the Eastern Cordillera. The timing of Eastern
Cordillera deformation derived from these cross
sections is late Miocene to Pliocene. However,
growth strata provide strong evidence of Late
Cretaceousearly Miocene Eastern Cordillera
deformation, as discussed later.
The Bucaramanga Fault along the western
sides of the Santander and Santa Marta Massifs
is a left-lateral strike-slip fault with a component
of west-verging, reverse movement (Fig. 1B).
Left lateral movement along this fault started
during the late Oligocene and has been accommodated by reverse faulting in the Santander
Massif and Eastern Cordillera (Toro, 1991).
Estimates of the total amount of sinistral displacement range between 100 and 115 km (e.g.,
Pindell et al., 1998); offset features include crystalline and Mesozoic rock units of the Central
Cordillera relative to the Santa Marta Massif
and the Cesar-Ranchera Basin relative to the
MMVB (Campbell, 1968; Fig. 1B).
CONSTRAINTS ON SEDIMENTARY
FILL EVOLUTION
This major section describes important aspects of the sedimentary fill of the
MMVB, Eastern Cordillera, and Llanos Basin
(Figs. DR1DR31), which are essential for our
reconstructions of basin configurations in later
sections of this paper.
1
GSA Data Repository item 2005148, Figures DR1
DR5 and Tables DR1DR3, is available on the Web at
http://www.geosociety.org/pubs/ft2005.htm. Requests
may also be sent to editing@geosociety.org.

Middle Magdalena Valley Basin


Here we synthesize conclusions of MMVB
studies by Gmez et al. (2003, 2005). An outstanding feature in the MMVB and adjacent
Eastern Cordillera foothills is the MMVU
(Figs. DR1A and DR1B, see footnote 1). This
surface bevels deformed pre-Eocene rocks
and is overlain by onlapping middle Eocene
to Neogene strata. Stratigraphic features of the
southern MMVB reveal that two long-lasting
events produced the MMVU (Fig. DR1A, see
footnote 1): (1) Late Cretaceousearly Eocene
eastward migrating Central Cordillera uplift,
and (2) consequent formation of a westwardexpanding pediment zone, a process still active
in the present Central Cordillera slope. Zircon
and apatite fission track data point to erosion
of 713 km of Central Cordillera rocks since
the Late Cretaceous, which has translated into
large sediment supply to the basins to the east
(Gmez et al., 2003, 2005).
In the southern MMVB, the record of Central Cordillera uplift is a Maastrichtian-Paleocene sequence of marine to alluvial fan facies
(Fig. DR1A, see footnote 1) with paleoflow and
petrography indicative of Central Cordillera
provenance. Plutonic granitic clasts appear in
the Paleocene Hoyn conglomerates and indicate early Tertiary erosion of deep levels of the
Central Cordillera. Northward propagation of
Central Cordillera uplift is recorded by a change
to continental facies and Central Cordillera
provenance, both of which occur in the Paleocene Lisama Formation of the northern MMVB
(Fig. DR1B, see footnote 1; Campbell, 1968).
MMVB Paleocene deposits were partially
eroded during continued early Eocene eastward
propagation of Central Cordillera uplift.
The Central Cordillera has been erosionally
beveled since middle Eocene times. Its boundary
with the MMVB has moved westward since then
as alluvial deposits onlapped over the residual
pediment (the MMVU, Fig. DR1A) and local
paleohighs (Fig. DR1B, see footnote 1). Eastern
Cordillera folding controlled middle Eocene
Neogene sedimentation and was diachronous
along the MMVB as indicated by the ages of
associated growth strata (Figs. 2A, 2B, DR1A,
and DR1B, see footnote 1). Specifically, middle
EoceneOligocene deformation of the Villeta
anticlinorium, east of the southern MMVB,
is indicated by growth strata in the lower and
middle parts of the San Jun de Ro Seco Formation (Figs. 2A, 2B, and DR1A, see footnote 1). In
the northern MMVB, younger deformation of the
Los Cobardes, Provincia, and Lisama anticlines
(Fig. 1B) is recorded by growth strata of late
Oligoceneearly middle Miocene age equivalent
to the upper part of the Mugrosa and Colorado

Geological Society of America Bulletin, September/October 2005

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

Figure 2. (A) Geologic map


of the southern Middle Magdalena Valley Basin (MMVB)
showing the locations of the
seismic lines in Figures 2B and
2C. (B) Portion of a depthconverted seismic line across
the Guaduas syncline. The
lower and middle San Jun de
Ro Seco synorogenic wedge is
made of smaller-scale syntectonic units, which are bounded
by growth unconformities
(GU). Overall divergence of
strata and lateral offsets of the
anticlinal axial surface along
the growth unconformities
were produced by pulses of
synsedimentary deformation
(e.g., Ford et al., 1997). The
overlying portion of the San
Jun de Ro Seco Formation
and the Santa Teresa Formation completely overlapped
this Eastern Cordillera uplift.
(B) Seismic line across the
southern MMVB west of the
Guaduas syncline. Strata of the
San Jun de Ro Seco Formation truncate against the base
of the Honda Group, and Neogene strata truncate against
the surface of the southern
MMVB. These configurations
were produced by basin rotation in response to loading
pulses from the Eastern Cordillera. See Gmez et al. (2003,
2005) for full description and
interpretation of MMVB seismic information.

Geological Society of America Bulletin, September/October 2005

1275

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

Formations (Fig. DR1B, see footnote 1). Three


Cenozoic fossil horizons record adjustments of
the MMVB alluvial plain profile (Fig. DR1, see
footnote 1), two of which correlate with eustatic
highstands (Los Corros and La CiraSanta Teresa
units, e.g., Haq et al., 1988).
Eastern Cordillera deformation also produced
regional-scale unconformities due to flexural
tilting. Two of these surfaces occur to the west
of the Eastern Cordillera thrust and fold belt.
The older is an unconformity at the base of
the Honda Group in the southern MMVB; the
youngest is the present surface of the MMVB
(Figs. 2A and 2C). Both surfaces merge with
the MMVU (i.e., slope of the Central Cordillera) and both truncate progressively older strata
toward the west. These unconformities indicate
eastward rotation of the basin in response to
episodes of increased tectonic loading from
the Eastern Cordillera combined with Central
Cordillera erosional unloading (Gmez et al.,
2003, 2005).
Eastern Cordillera
The Upper CretaceousPaleogene Eastern
Cordillera sedimentary record is composed
of three sequences in the Bogot and Boyac
regions (Fig. DR2, see footnote 1). The first
sequence, Upper Cretaceouslower Paleocene
Guadalupe Group and Guaduas Formation,
shallows upward, recording sea withdrawal
from this region (Fabre, 1983b; Sarmiento,
1994). The Guadalupe-Guaduas sequence is
overlain by two fining-upward sequences of
upper Paleocenelower Eocene and middle
EoceneOligocene ages, both with fluvial
sandstones at their bases (Cuervo and Ramrez,
1985; Acosta and Beltrn, 1987). Nonmarine
conditions were firmly established in the Sabana
de Bogot area during the second sequence but
marine influence increased toward the north
(Fig. DR2, see footnote 1). In the third sequence
(middle EoceneOligocene), fluvial sandstones
of the Regadera Formation (Hoorn, 1988) and
equivalent Picacho Formation are transitional
upward into mudstones with foraminifera and
oolitic iron of the Usme and Concentracin
Formations, which record renewed marine
influence in the Eastern Cordillera (Hubach,
1957; Acosta and Beltrn, 1987). The Eastern
Cordillera Paleogene rocks are unconformably
overlain by upper Miocene to Pliocene alluvial
and lacustrine deposits (Marichuela and Tilat
Formations). They formed in intramontane settings during the most intense phase of Eastern
Cordillera uplift (Helmens, 1990).
The Usme syncline provides detailed information to reconstruct Late CretaceousPaleogene relations typical of Eastern Cordillera

1276

Figure 3. Upper CretaceousOligocene growth strata of the Usme syncline. (A) Geometry of
the Usme syncline extracted from structural data provided by Julivert (1963). (B) Detailed
mapping of the eastern flank of the Usme syncline by Julivert (1963) reveals the occurrence
of unconformities. (C) Reconstruction of the geometry of strata and unconformities in
Figure 3B results in a westward-thickening wedge of strata with internal angular unconformities, which become conformable surfaces toward the west. Upper Cretaceouslower
Oligocene strata are thicker and conformable at the core of this syncline. (D) The geometries
of growth strata of the Pyrenees associated with progressive limb rotation (Ford et al., 1997)
are similar to the stratal configurations found in the Usme syncline and MMVB.

folds (Figs. 1B and 3; Julivert, 1963). Upper


CretaceousOligocene strata are conformable at
the core of this syncline but become thinner and
unconformable toward the eastern flank, where
angular relations locally reach 90. In addition

to an absence of Maastrichtian beds (Gp4 and


Gs in Figs. 3B and 3C) on the eastern flank, several other unconformities occur within the overlying Paleogene section. We interpret that these
geometries resulted from progressive rotation of

Geological Society of America Bulletin, September/October 2005

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

Figure 4. Paleontologic age assignments of Upper CretaceousPaleogene units of the


Eastern Cordillera are overlapping in the Bogot and Boyac (names within parentheses)
regions. Therefore, there is no chronological basis to postulate the occurrence of a regional
unconformity equivalent to the MMVU in central areas of the Eastern Cordillera. Wider
bars indicate our preferred interpretation of depositional ages. The palynologic ages of
Paleogene units of the Boyac region (gray bars) are broadly similar to the ages of their
equivalents in the Bogot area (black bars). Most important sources of ages: Hubach (1957),
Van der Hammen (1957), Germeraad et al. (1968), Acosta and Beltrn (1987), Hoorn (1988),
Helmens (1990), and Cspedes and Pea (1995).

the eastern limb of the Usme syncline, as in the


case of similar syntectonic strata found in the
MMVB (Fig. 2B) and in the Pyrenees (Ford et
al., 1997, Fig. 3D herein). Julivert (1963), Raasveldt (1956), and Laverde (1989) also document
variable patterns of unconformities, thicknesses,
and facies in other folds located south of the
Bogot region and along the Upper Magdalena
Valley Basin, which are evidence of broader
Paleogene Eastern Cordillera synsedimentary
deformation.
In contrast with the MMVB, there are no
physical or chronological bases to postulate the
occurrence of an unconformity equivalent to the
MMVU in central areas of the Eastern Cordillera, as previous works did (e.g., Cooper et al.,
1995; Pindell et al., 1998; Villamil, 1999). The
vertical bars in Figure 4 describe the maximum
time duration of each Maastrichtian to Cenozoic
stratigraphic unit of the Eastern Cordillera permissible by data. Maastrichtianearly Paleocene
ages are based on ammonites, foraminifera, and
pollen (Etayo-Serna, 1964, 1985; Sarmiento,
1994). The age assignments of the overlying
Paleogene units rely on palynology and are
coarse due to incomplete palynologic sampling
(Hubach, 1957; Van der Hammen, 1957; Hoorn,
1988), owing in part to the occurrence of growth
unconformities at sampling sites (e.g., eastern
flank of the Usme syncline; Hoorn, 1988).
Despite incomplete palynologic sampling, the
ages of successive Paleogene units overlap

(Fig. 4), which indicates that major-rank unconformities do not exist in the Sabana de Bogot
region. In contrast, the duration of the time
gap associated with the MMVU in the Eastern
Cordillera western foothills is on the order of
1015 m.y. (Fig. DR1, see footnote 1).
The preservation of Upper CretaceousOligocene Eastern Cordillera growth strata (Fig. 3)
required continuous accumulation of sediment
over a time scale (~1 m.y.) substantially less
than the total duration of each unit (Fig. 4; e.g.,
Crowley, 1984; Anders et al., 1987; Anadn et
al., 1986; Ford et al., 1997). We emphasize that
although minor temporal gaps exist within Upper
CretaceousPaleogene strata flanking Eastern
Cordillera folds, the mechanisms that produced
these unconformities were associated with local
synsedimentary folding rather than with regional
uplift and pedimentation of the Central Cordillera, as in the case of the MMVU.
Llanos Foothills and Llanos Basin
The Late CretaceousCenozoic stratigraphy
of the Llanos foothills and Llanos Basin has
been comprehensively studied by Cooper et
al. (1995; Fig. DR3, see footnote 1). The main
sediment source areas of these regions were the
Guyana Shield, during the Late Cretaceous to
early middle Miocene, and the Eastern Cordillera since the late middle Miocene (Cooper
et al., 1995). The MMVU unconformity of

the MMVB correlates with two major timetransgressive unconformities in the Llanos
foothills, which merge into a single composite
unconformity further east in the Llanos Basin
(Fabre, 1983b; Cooper et al., 1995). The older
surface truncates deeper levels of the Guaduas
Formation and Guadalupe Group from west to
east (Sarmiento, 1994). The Guaduas Formation and upper Guadalupe Group are absent
in the Llanos Basin, where upper Paleocene
sandstone-rich stuarine and coastal plain mudstones of the Barco and Los Cuervos overlie
Campanian rocks (Cooper et al., 1995). A second major unconformity separates these units
from the overlying upper Eocene to Neogene
eastward onlapping megasequence. The upper
Eocenelower Miocene strata of the Llanos
Basin consist of alluvial plain, coastal plain, and
estuarine valley-fill deposits; the continental
character of these deposits increases toward the
east. Sedimentation initiated with deposition of
the Mirador Formation sandstones and continued with deposition of four eastward stepping,
coarsening upward sequences in the Carbonera Formation (Fig. DR3, see footnote 1).
The Carbonera is capped by the Len Shale,
which marks a major marine transgression.
Major Eastern Cordillera uplift is indicated by
provenance of the middle MiocenePliocene
Guayabo Formation (Cooper et al., 1995).
GEOHISTORY ANALYSIS OF THE
MMVB AND EASTERN CORDILLERA
As a first step to extract the signal of tectonic
subsidence to the east of the Central Cordillera, we carried out one-dimensional geohistory analyses of three columns exposed in the
southernmost MMVB (Villeta anticlinorium
and Guaduas syncline), northern MMVB (Los
Cobardes anticline and Nuevo Mundo syncline), and the Cocuy anticlinorium (Boyac
region, Figs. 1B, 5A5C; Fabre, 1983b, 1985;
Sarmiento, 1989; Gmez, 2001). The backstripping techniques are described by Allen and
Allen (1992). The sedimentary columns were
decompacted assuming a depth-dependent,
exponentially decreasing porosity function. In
order to estimate the tectonic subsidence, the
subsidence driven by the load of sediment was
subtracted from the total (decompacted) subsidence; corrections for varying water depth and
long-term eustatic sea level (Haq et al., 1988)
were applied. Summary tables of stratigraphic
attributes and physical parameters used for
backstripping are available (Tables DR1DR3,
see footnote 1). Constraints on Tertiary paleoaltitudes are poor (see error bars in Fig. 6). We
have assumed an average altitude between 0 m
and 500 m for the MMVB, which corresponds

Geological Society of America Bulletin, September/October 2005

1277

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.
Eustatic Sea Level (m)
250

N MMVB

0
2

6
8

uplift

Depth (km)

10
12

14
16

thermal
subsidence
(TS)

synrift
subsidence
(SS)

200 180 160 140 120 100

80

60

40

Figure 5. Geohistory plots of the MMVB and


Eastern Cordillera extracted from backstripping of sedimentary columns. (A) Northern
MMVB (Los Cobardes anticline and Nuevo
Mundo syncline). (B) Southern MMVB (Villeta anticlinorium and Guaduas syncline).
(C) Cocuy region of the Eastern Cordillera,
modified after Fabre (1983b). See Figure 1B
for locations. Facies, ages, and water depths
of Cretaceous units from Sarmiento (1989),
Fabre (1985), and Etayo-Serna (personal
commun., 2000). Late CretaceousMesozoic
parts of the subsidence curves are enlarged
in Figure 6, with error bars. Curves of subsidence corrected for sediment load nearly
overlap the final tectonic subsidence curves
because the correction for long-term eustatic
change is small.

20

Age (Ma)
Jurassic

Paleogene Neogene

Cretaceous

Eustatic Sea Level (m)


250

S MMVB

B
Depth (km)

Mesozoic Subsidence

uplift

6
8
10

synrift
subsidence

12
140 120 100

TS
80

60

40

20

Age (Ma)
Paleogene Neogene

Cretaceous

Eustatic Sea Level (m)


250

Cocuy

Depth (km)

2
4
6

thermal

SS
subsidence
8
140 120 100 80 60

40

Age (Ma)
Cretaceous

Paleogene

compacted rock thickness


decompacted
(total subsidence)
sediment load corrected
tectonic subsidence

1278

to the present altitude limits of the MMVB


alluvial plain. Facies in the Cocuy region and
Llanos Basin reflect eustatic influence and indicate lower altitudes.

part of subsidence
explained by load
of strata

The backstripped sections illustrate the subsidence history of the Mesozoic MagdalenaTablazo, Cocuy, and Cundinamarca extensional
sub-basins (Fig. 1B). Mesozoic synrift subsidence in these localities is independently recorded
by unconformity-bounded units, rotated-block
morphology, spatially variable thicknesses,
bimodal subaerial volcanogenic strata, and mafic
intrusives (Julivert, 1958; Fabre and Delaloye,
1982). Lithospheric extension is expressed on the
backstripped curves as rapid and temporally variable tectonic subsidence (Fig. 5). Subsequently,
smooth decline of the subsidence curves indicates postrift thermal contraction during lithospheric cooling (Allen and Allen, 1992).
The synrift tectonic subsidence of the northern MMVB lasted until the end of the Jurassic
(2993 m, 59 m/m.y., Fig. 5A). Synrift subsidence in the Villeta anticlinorium (Cundinamarca
sub-basin) until the Coniacian is indicated by
the stair-shaped synrift tectonic subsidence
until 87 Ma (2892 m, 56 m/m.y. average rate,
Fig. 5B). This history correlates well with
Hauterivian to Coniacian gabbroic sills of the
northern part of the Villeta anticlinorium (Fabre
and Delaloye, 1982; Rodrguez and Ulloa,
1994a, 1994b), which attest to extension and
injection of mafic magmas due to partial adiabatic melting (Turcotte and Schubert, 1982).
This prolonged synrift subsidence explains the
accumulation of ~7 km of Cretaceous strata in

Geological Society of America Bulletin, September/October 2005

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

80

Age (Ma)
40

Depth (km)

20

Umir

Lis

TS

FS

MMVU Paz

Esm

Eustatic Sea Level Curve 250 m


0m

Mug Col Real

Northern MMVB
FS

Umir: Umir Formation


Lis: Lisama Formation
Paz: La Paz Formation
Esm: Esmeraldas Formation
Mug: Mugrosa Formation
Col: Colorado Formation
Real: Real Group

3
4

Cretaceous Paleocene Eocene Oligocene Miocene

80

60

Age (Ma)
40

20

Eustatic Sea Level Curve 250 m

Ol

0m
Southern MMVB

Buscav Se Hoy MMVU Arm Alm Cruz ST


+Cim

Ol: Olini Group


Buscav: Buscavidas Shale
Cim: Cimarrona Formation
Se: Seca Formation
Hoy: Hoy n Formation
San Ju n de R o Seco Formation:
Arm: Armadillos member
Alm: Alm cigos member
Cruz: La Cruz member
ST: Santa Teresa Formation

3
4
TS

Late Cretaceous to Neogene Subsidence

FS

FS

Cretaceous Paleocene Eocene Oligocene Miocene

80

60

Age (Ma)
40

20

Eustatic Sea Level Curve 250 m

C
L Pin T G

Depth (km)

The Late CretaceousCenozoic subsidence


curves (Figs. 6A6C) highlight two aspects of
regional accommodation. First, continuous subsidence characterized the MaastrichtianPaleocene and middle EoceneNeogene times in the
MMVB and the Late CretaceousOligocene
in the Cocuy region. Tectonic subsidence was
larger than the contemporaneous oscillations
of eustatic sea level. The resultant long-term
high accommodation explains the preservation of abundant muddy deposits in the Upper
CretaceousCenozoic sedimentary record. For
the two MMVB sites, the intervals of subsidence are separated by a time of uplift. Second,
sediment loading was the most important force
driving total subsidence and amounts for ~70%
of the total Tertiary subsidence, as illustrated
by the difference between total subsidence and
tectonic subsidence in Figure 5.
Independent of manipulation of error bars in
the subsidence plots, tectonic subsidence rates
seem to have increased during the late MaastrichtianPaleocene in the southern MMVB
(Fig. 6B, 433 m, 37 m/m.y.) and during the
Paleocene in the northern MMVB (Fig. 6A,
290 m, 32 m/m.y.) relative to the Cretaceous
thermal subsidence. The tectonic subsidence
curve of the Cocuy region (Fig. 6C, 413 m,
11 m/m.y.) does not show this behavior. The
increased rates of tectonic subsidence of the

60

Depth (km)

the Villeta anticlinorium area, the thickest in the


Eastern Cordillera. The synrift subsidence in
the Eastern Cordillera Cocuy region (1960 m,
73 m/m.y.) spans the Berriasian to Aptian time
interval (144117 Ma, Fig. 5C), according
to subsidence history and mafic magmatism
(Fabre, 1983b).
Thermal sagging spanned the Cretaceous in
the northern MMVB (933 m, 12 m/m.y.), the
late ConiacianMaastrichtian in the southern
MMVB (195 m, 10 m/m.y.), and the Aptian
Maastrichtian in the Cocuy region (893 m,
18 m/m.y.; Figs. 5A5C). Thermal subsidence
curves are suggestive of stretching factors ()
of 1.4, 1.2, and 1.6, respectively (Gmez, 2001).
However, these values most likely underestimate , given the heat loss during the long
rifting (Allen and Allen, 1992). Previously, it
was suggested that thermal sagging of a passive
margin produced subsidence east of the Central
Cordillera during most of the Cretaceous (Pindell, 1993; Roeder and Chamberlain, 1995).
However, the temporally and spatially variable
subsidence histories described in this section
most likely correspond to back-arc subsidence
behind a magmatic arc in the Central Cordillera
(e.g., Aspden et al., 1987; Cooper et al., 1995).

L US Pic
S

0m
Cocuy region, E. Cordillera

Conc

L: La Luna Formation
Pin: Los Pinos Formation
T: Tierna sandstone
Pin: Los Pinos Formation
G: Guaduas Formation
LS: Lower Socha Formation
US: Upper Socha Formation
Pic: Picacho Formation
Conc: Concentraci n Formation

3
4

TS

FS

Cretaceous Paleocene Eocene Oligocene Miocene


TS: thermal subsidence
FS: flexural subsidence

MMVU: Middle Magdalena Valley unconformity


period of unconformity formation

Figure 6. Late Cretaceousearly Miocene tectonic subsidence of the (A) northern MMVB,
(B) southern MMVB (Gmez, 2001), and (C) Eastern Cordillera (Cocuy region, modified
after Fabre, 1983b). Vertical and horizontal error bars represent uncertainties in paleoaltitude and stratigraphic ages, respectively. See text for discussion.
southern and northern MMVB correlate with
Central Cordillera uplift and appear to support interpretations of foreland basin subsidence triggered by Central Cordillera loading
(Dengo and Covey, 1993). Because tectonic

subsidence rates were much higher than the


rate of contemporaneous long-term sea-level
drop (1 m/m.y., Haq et al., 1988), one might
suppose that the basin would have remained
flooded by marine waters. The regressive facies

Geological Society of America Bulletin, September/October 2005

1279

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

NM: Nuevo Mundo syncline


LC: Los Cobardes anticline
AA: Arcabuco anticline
B-C: Boyac -Cocuy outcrops
GS: Guaduas syncline
VA: Villeta anticlinorium
FS: Fusagasug syncline
US: Usme syncline
LLF: Llanos foothills outcrops

NM

LC

AA
B-C

GS
VA

LLF
US
FS

Figure 7. Early Paleocene palinspastic base map constructed by combining Dengo and
Coveys (1993) and Dazs (1994) reconstructions with restored sections from other authors
(numbered lines). This retrodeformation, perpendicular to the Eastern Cordillera structural
strike, also restores the continuity between stratigraphic units of the Central Cordillera and
Santa Marta Massif and between the MMVB and the Cesar-Ranchera basin across the
Bucaramanga fault. See text for discussion and other sources of data. Features marked as
reference outcrops form a template visible in the paleogeographic maps of Figures 8, 10,
1517, DR4, and DR5 (see footnote 1). See also Figure 1B for present location of outcrops.

(Fig. DR1, see footnote 1) clarify that this was


not true. Thus, under conditions of enhanced
accommodation relative to Late Cretaceous
times, the MMVB MaastrichtianPaleocene
sea withdrawal is explained by the large input
of sediment from the Central Cordillera. The
combined enhanced tectonic subsidence and
facies patterns refute the interpretation that
Paleocene regression was caused by decreasing
accommodation resulting from Central Cordillera deformation combined with the drop of
eustatic sea level (e.g., Villamil, 1999).
Early Eocene kilometer-scale uplift in the
MMVB is indicated by apatite-fission-track
parameters from both the Central Cordillera
basement and Mesozoiclower Paleogene
sedimentary rocks of the MMVB (Gmez
et al., 2003, 2005). Middle Eocene to early
middle Miocene subsidence rates of the southern MMVB (468 m, 15 m/m.y.) and northern
MMVB (649 m, 21 m/m.y.) are lower than
MaastrichtianPaleocene values.

1280

RECONSTRUCTION OF LATE
CRETACEOUSCENOZOIC BASIN
GEOMETRIES AND HYPOTHESES
OF MECHANISMS OF LONG-TERM
SUBSIDENCE
In this section, we describe the construction
of a palinspastically restored base (Fig. 7),
which serves as a template for reconstruction
of two major basin configurations of Late Cretaceousearly Eocene and middle EoceneNeogene ages (Figs. 811). We also correlate the
previously summarized information from the
MMVB, Eastern Cordillera, and Llanos Basin
along the cross sections in Figures 9 and 11.
Palinspastic Reconstruction
As previously described, deformation
has modified the original geometries of the
Colombian basins, and therefore palinspastic
reconstruction is required to reconstruct the

wavelength of regional subsidence. Retrodeformation perpendicular to the Eastern


Cordillera structural grain, a condition of pure
shear, should also restore the original continuity between the Central Cordillera and the
Santa Marta Massif and between the northern
MMVB and the Cesar-Ranchera Basin across
the Bucaramanga fault (e.g., Fig. 1B). This
constraint favors structural models that predict
at least 150 km of Eastern Cordillera crustal
shortening (e.g., Dengo and Covey, 1993;
Roeder and Chamberlain, 1995). Dengo and
Coveys (1993) cross section (Fig. 1C) is well
suited for restoration because they describe
their supporting data and validate their section
with a gravity profile. Their restoration of Eastern Cordillera shortening also best restores the
~110 km of sinistral offset along the Bucaramanga fault. We used this restoration to move
the Paleogene and Jurassic outcrops of the
northern part of the Eastern Cordillera and
the Central CordilleraMMVB block to their
position in the early Eocene (Fig. 7). Retrodeformation generates an opening between the
present trace of the Bucaramanga fault and the
northern MMVB, whose partial filling requires
counterclockwise rotation of the Santander and
Santa Marta Massifs by ~12. The remaining
unfilled space may represent compressional
deformation in the MMVB and Santa Marta
Massif. A long-term history of pedimentation
indicates that no important Central Cordillera shortening has occurred since the middle
Eocene (Gmez et al., 2003). Thus, we treat
the Central Cordillera and MMVB as a single
rigid block, which is transported toward the
NW during restoration. No estimates of Late
Cretaceousearly Eocene Central Cordillera
shortening are available. Thus, no further
retrodeformation of this range is attempted.
The Llanos Basin area, overlying the Guyana
Shield, is assumed to remain fixed during
restoration. The palinspastically restored base
of Daz (1994), based on numerous balanced
cross sections, was used to restore the southern
portion of the Eastern Cordillera at the latitude
of Bogot. Restoration of the Eastern Cordillera foothills is also constrained by cross sections across the MMVB, Llanos foothills, and
the Perij range (Fig. 7).
Palinspastic restoration is an exercise with
considerable assumptions and error. Simple
shear conditions may lead to more accurate
restored maps, but they require oblique fault
displacement and rotation of thrust slices, which
are unconstrained. Our restoration is simple,
meets major regional constraints, and offers
a better approximation of pre-Neogene basin
dimensions for the subsequent basin analyses
than do modern spatial relations.

Geological Society of America Bulletin, September/October 2005

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

Late Cretaceous to Early Eocene Basin


Geometry and Hypothetical Causes
The restored isopach map and cross section
of Maastrichtianlower Eocene strata to the east
of the Central Cordillera reveal an asymmetric
distribution of sedimentary thicknesses (Fig. 8).
The reconstructions of geometric and temporal
relations among strata are shown by the restored
cross section and accompanying chronostratigraphic diagram in Figures 9A and 9B. Thicknesses decrease toward the east from the Eastern
Cordillera to the Llanos Basin (Figs. 8 and 9A).
Maastrichtian to Paleocene rocks in the southern
MMVB have a maximum thickness of 1897 m.
However, they were truncated beneath the
MMVU, and their initial total thickness was
greater; thermal history parameters indicate
erosion of a lower Paleogene sedimentary section up to 3 km thick in the Guaduas syncline
(Gmez et al., 2003). The maximum thickness
preserved of the Maastrichtianlower Eocene
section is in the western side of the Sabana de
Bogot region (~2400 m). Diminished thicknesses toward the east resulted from diminished
eastward accommodation and the presence of
two regional unconformities. The thickness of
Maastrichtianlower Eocene strata in the Cocuy
region is 1109 m (Fabre, 1985) and diminishes
to zero toward the Llanos Basin region. Geohistory analyses suggest that Maastrichtianearly
Eocene tectonic subsidence decreased eastward
from 500 m in the southern MMVB to 205 m in
the Cocuy region (Fig. 9A, inset). The depositional topographic slope likely decreased from
alluvial-fan gradients (>2) in the west to lower
alluvial- and coastal-plain gradients in the Eastern Cordillera and Llanos Basin (<0.4, e.g.,
Blair and McPherson, 1994; Daz, 1994).
The wedge-shaped sedimentary fill in
Figure 9A resembles the typical geometry of
a foreland basin, where isostatic subsidence is
distributed laterally by lithospheric flexure across
distances of hundreds of kilometers from the
mountain belt. The details of wavelength and
amplitude depend on the flexural rigidity of the
lithosphere (Jordan, 1981, 1995). Flexural subsidence diminishes to zero on the distal margin,
where there may be a flanking peripheral bulge. In
an ideal foreland basin, the mountain front moves
toward the foreland with a self-similar geometry,
and foreland basin subsidence and fill migrate
simultaneously with it. Periods of active deformation and tectonic quiescence produce variations
in the basins flexural profile. Enhanced tectonic
thickening in the mountain belt produces a basin
that is relatively narrow and deep, with enhanced
preservation potential of fine-grained facies. At
those times, the marginal bulge becomes a site
of erosion and sediment bypass. During tectonic

Figure 8. (A) Palinspastically restored distribution of preserved thicknesses of Maastrichtian


lower Eocene strata to the east of the Central Cordillera. Outlines of the Central Cordillera,
Santa Marta Massif, restored positions of the Eastern Cordillera boundaries, and some
Eastern Cordillera outcrops are provided for reference (e.g., Figs. 1B and 2). (B) Cross section displaying a large vertical exaggeration (5) to illustrate thickness variation of strata. A
much larger vertical exaggeration is needed to represent internal distributions of facies and
unconformities (see Fig. 9A).

quiescence, surface processes strip material from


the mountain and transfer it to the basin, resulting
in isostatic rebound of the source area. But the
lack of tectonic subsidence in the basin causes
accumulation over a wider zone, which results
in depression of the peripheral bulge. Sequence
boundaries cap the quiescent-phase strata (Flemings and Jordan, 1990).
If a Maastrichtianearly Eocene foreland
basin existed east of the Central Cordillera,
then the two Paleogene regional unconformi-

ties in the eastern sector (Llanos region) may


point to alternating periods of Central Cordillera eastward-progressing uplift and tectonic
quiescence. The distribution of the regional
unconformities in Figures 9A and 9B indicate
that the basin wavelength decreased from more
than 320 km during the Maastrichtian to 150 km
during the Danian and from more than 350 km
in the latest Paleocene to 150 km during the
early Eocene. We hypothesize that the phase
of basin widening to 350 km correlates with

Geological Society of America Bulletin, September/October 2005

1281

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

Figure 9. (A) Palinspastically


restored cross-section configuration of the Maastrichtian to
early Eocene foreland basin
adjacent to the Central Cordillera. Vertical exaggeration
~65. (B) Corresponding chronostratigraphic section. See text
for discussion.

late Paleocene tectonic quiescence, isostatic


rebound of the Central Cordillera, and ensuing depression of the Llanos peripheral bulge.
Similarly, the Danian and early Eocene stages
of basin narrowing may reflect times of Central
Cordillera shortening and uplift and westward
displacement of sedimentation to the locus of
maximum subsidence. The area of maximum
thickness in the Sabana de Bogot region marks
the final early Eocene position of the zone of
maximum accommodation. Continuous sedimentation characterized this region, although
the foreland basin was disrupted by small
synsedimentary folds (e.g., Usme Syncline)
that produced growth unconformities of limited
extent (not shown in Fig. 9).
Middle Eocene to Neogene Basin Geometry
and Hypothetical Causes
Regional deformation at the latitude of the
MMVB changed from the Central Cordillera
to the Eastern Cordillera during the middle
EoceneNeogene. Pedimentation was the dominant process in the Central Cordillera, whereas
widespread Eastern Cordillera deformation is
revealed by thermal-history and provenance data
of the MMVB and by growth strata flanking the
Usme and Fusagasug synclines and the Villeta,
Provincia, and Lisama anticlines (Fig. 1B;
Gmez et al., 2003, 2005). Basin geometry also

1282

differed during the middle Eocene to Neogene,


compared to the previous patterns, as shown
by the distribution of thicknesses of units of
these ages (Fig. 10). These strata thin regionally
toward both the west and east due to sedimentary onlap onto the Central Cordillera pediment
surface and the Guyana Shield (e.g., Figs. 10,
DR1, and DR3, see footnote 1). Maximum
preserved thicknesses of middle EoceneNeogene strata are in the MMVB (~7000 m) and
Llanos foothills area (~4500 m). Stratigraphic
thicknesses are not well constrained in the
Eastern Cordillera area (gray-shadowed area in
Fig. 10A). Scarce outcrops of these ages in this
region have maximum preserved thicknesses on
the order of ~1500 to ~1700 m.
The cross sections in Figure 11 synthesize
our interpretation of middle EoceneOligocene
basin geometry. Eastern Cordillera anticlines
segmented the region to the east of the Central
Cordillera, as revealed by growth strata in the
Usme and Guaduas synclines. Conglomeratic piedmont facies indicate a high-gradient
topographic profile in the southern MMVB,
which decreased eastward across the synclinal
basins of the Eastern Cordillera to coastal-plain
gradients in the Llanos Basin. The gradient of
the residual slope of the Central Cordillera
(MMVU) was persistently steeper than the topographic gradient of the Guyana Shield to the
east, which resulted in faster onlap further east.

There is no evidence that Central Cordillera


tectonic thickening in the middle EoceneNeogene caused basinal subsidence. Unlike during
the Maastrichtianearly Eocene, during this
later time the MMVB strata passively onlap
westward onto the Central Cordillera, suggesting progressive decrease in its elevation
(Gmez et al., 2003). Erosional denudation of
the Central Cordillera would have contributed
flexural uplift to the MMVB rather than subsidence. Isostatic adjustment to erosion explains
the net decrease of Central Cordillera altitude.
The explanation for tectonic subsidence in the
MMVB and Llanos Basin rests ultimately with
Eastern Cordillera thickening. A small amount
of tectonic subsidence was amplified into a
larger amount of total subsidence (e.g., Fig. 5)
because sediment was trapped between the Central Cordillera and uplifts in the Eastern Cordillera and between these uplifts and the Guyana
Shield (Fig. 11). Onlap of these sediments
toward the west and east caused additional
loading and flexural subsidence. Sediment
compaction also contributed a modest amount
of space (~6% of total subsidence according to
backstripping) and increased toward the east in
the MMVB and toward the west in the Llanos
Basin because the sedimentary fill was thicker
in those directions.
The contribution of Eastern Cordillera tectonic loading to subsidence increased through

Geological Society of America Bulletin, September/October 2005

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

time as this range evolved into a continuous


and wide topographic feature. For example,
Neogene pulses of Eastern Cordillera uplift
have clear manifestations in the MMVB where
Cenozoic strata are regionally tilted toward the
east (e.g., Fig. 2C). The mechanisms of MMVB
Neogene tilting and erosional truncation of
strata are similar to those that generated Paleogene unconformities in the Llanos Basin area,
the difference being that the Llanos Paleogene
surfaces were produced by flexure under Central Cordillera loading.

MECHANICAL MODELING: TESTING


HYPOTHESES OF BASIN SUBSIDENCE
In this section, we test the mechanical viability of our interpretations of subsidence mechanisms. The simplest test is to compute in two
dimensions the flexural deformation produced
by Central and Eastern Cordillera crustal loads
on an infinite, elastic lithosphere of constant
thickness (Turcotte and Schubert, 1982). If the
mechanical controls are as previously hypothesized and the geologic constraints are accurate,
this modeling should reproduce the basin configurations that were interpreted from geologic
data; failure of the mechanical models to fit the
observations would invalidate our hypotheses.
Numerical experiments were conducted
according to procedures described by Cardozo and Jordan (2001). The two-dimensional
modeling strategy involves the conversion of
tectonic and sedimentary loads into rectangles
of equal width (w), but different height (hi) and
density (i) (Fig. 12). The deflection profile
[ui(x)] of each one of these elements (i g hi)
is then computed based on differential equations
that describe the flexure of an elastic lithosphere
(Turcotte and Schubert, 1982). The total displacement profile [u(x)] is equal to the sum of
all the deflection profiles of the rectangular elements. The sum of the load profile [h(x)] and the
displacement profile [u(x)] is the relative topographic profile [rt(x)], which displays elevation
relative to an initially undeformed reference
datum (Fig. 12).
We selected four cases to simulate the flexural response to crustal loads of the Central and
Eastern Cordilleras. Discrete sedimentary loads
(w = 10 km, Fig. 13) were extracted from the
cross sections in Figures 9A and 11A. Discrete
sedimentary thicknesses were decompacted by
a factor of 6%, an average extracted from backstripping. In all cases, we evaluated the fit of the
model approximations of paleotopography by
comparing it to the facies pattern in the basin.
The first three experiments (Maastrichtian,
early, and late Paleocene configurations) test
the Late Cretaceousearly Eocene foreland

VA
BA

VA: Villeta anticline

VA

BA: Bogot anticline

BA

Figure 10. (A) Palinspastically restored distribution of thicknesses of middle EoceneNeogene


strata in the MMVB, Eastern Cordillera, and Llanos Basin. See text for discussion. Outlines
of the Central Cordillera, Santa Marta Massif, restored positions of the Eastern Cordillera
boundaries, and some Eastern Cordillera outcrops provided for reference (e.g., Figs. 1B
and 2). (B) Cross-section view of stratal-thickness variation. Vertical exaggeration, 2.

basin interpretation. This hypothesis relies on


the assumption that there was crustal thickening of the Central Cordillera. But because there
are not accurate structural data of this range
demonstrating the amount of crustal shortening,
the paleoaltitude of the Central Cordillera is the
best record of crustal thickness changes. In our
experiments, the Central Cordillera tectonic
load profiles are the best choices after several

iterations. As discussed earlier, kilometer-scale


paleoaltitudes of the Central Cordillera can be
inferred from its remnant topography, its degree
of denudation as revealed by thermal history
parameters of basement rocks, the early unroofing of Mesozoic granitic plutons revealed by
provenance data of lower Tertiary conglomerates, and effects on MMVB facies (Gmez et
al., 2003, 2005). Iterative testing showed that

Geological Society of America Bulletin, September/October 2005

1283

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

the best choice of lithospheric elastic thickness


(t) for modeling was 35 km. This value is intermediate between elastic thicknesses associated
with rifting (030 km) and cold continental
lithospheres (4590 km), which represent
extreme values for the Colombian lithosphere
during Mesozoic and late Tertiary times, respectively (Turcotte and Schubert, 1982; Allen and
Allen, 1992; see also Kellogg et al., 1995;
Roeder and Chamberlain, 1995).
Modeling Results

Figure 11. Palinspastically restored cross-section configuration of (A) late EoceneOligocene and (B) late Oligocene Colombian basins to the east of the Central Cordillera. Longterm subsidence was produced by loading from Eastern Cordillera folds and onlapping
sediments. See text for explanation.

Figure 12. Conceptual description of our two-dimensional mechanical models after Cardozo
and Jordan (2001). Loading of an elastic lithosphere overlying a semifluid asthenosphere
(left) is solved analytically using the analogy of an infinite beam on an elastic foundation
(middle and right). See text and Table 1 for explanation of constants and variables. Dark
and light gray denote mountain and sedimentary material, respectively.

1284

The three first experiments evaluate the flexural responses to uplift and tectonic quiescence of
the Central Cordillera (Figs. 13 and 14). Experiments 1 and 2 (Figs. 14A and 14B) simulate
Maastrichtianearly Paleocene eastward-advancing uplift of this range. The third experiment
simulates the redistribution of crustal loads
associated with late Paleocene Central Cordillera tectonic quiescence and erosion (Fig. 14C).
The reference datum is a middle Maastrichtian
surface at the bases of the Cimarrona Formation
and Tierna sandstone (Fig. 9A).
Our first model (Fig. 14A) evaluates the
Maastrichtian configuration beneath the La
Seca and Guaduas Formations (Fig. 9A). The
maximum modeled Central Cordillera paleoaltitude is 2000 m. The depositional profile of this
model is steeper on the western side of the basin
and gentler in the eastern side, as predicted by
piedmont and coastal facies, respectively. A
topographic low between 100 and 200 km in the
horizontal distance axis of Figure 14A is consistent with the locus of sedimentation of shallow
marine mudstones (Daz, 1994). Most of the
accommodation space is produced by flexure,
with a maximum thickness of 250 m being
accommodated above the undeformed reference
datum in the western side of the basin.
In the second experiment (early Paleocene,
Fig. 14B), we increased the height of the Central Cordillera loads and moved the boundary
between this range and the sedimentary basin
toward the east (e.g., Fig. 9). Total subsidence
next to the Central Cordillera is amplified
more than two times relative to Maastrichtian
subsidence (compare Figs. 14A and 14B). The
modeled position of the forebulge is displaced
toward the west relative to its Maastrichtian
position, which is consistent with the formation of a Late CretaceousDanian unconformity
along the eastern side of the basin. The continuation of this surface in the modeled basin
is delineated by the dashed line in Figure 14B,
representing the deflected position of what had
been the Maastrichtian topography. The flat
topography of the eastern side of the basin,
east of 320 km in the horizontal-distance axis

Geological Society of America Bulletin, September/October 2005

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

of Figure 14B, truncates westward-tilted Maastrichtianearliest Paleocene beds. Toward the


west, the slope of the depositional profile is the
steepest in the region adjacent to the Central
Cordillera, where alluvial fan deposition is
documented. This model suggests that a total
subsidence of 1700 m and paleoelevations on
the order of 300 m next to the Central Cordillera
provided enough space to accommodate most of
the lower Paleocene alluvial sediments sourced
by the Central Cordillera (Fig. 14B). Erosion
and sediment bypass dominated in the eastern
sector of the basin. Paleocene paleoflow in the
Sabana de Bogot region was oriented toward
the NNE (Laverde, 1989), which reflects the
effects of a northward slope gradient and intrabasinal folding. These features are ignored in
these two-dimensional models.
In the third experiment (Fig. 14C), we reduced
the height of the Central Cordillera loads to
simulate late Paleocene tectonic quiescence
and erosion. The boundary between Central
Cordillera and the foreland basin is kept at the
same position as in the early Paleocene model.
The interpreted loads of quiescent-phase strata
correspond approximately to the Cacho and
lower Bogot and the Barco and Los Cuervos
Formations (Fig. 9). Erosion produces isostatic
rebound of the Central Cordillera and proximal
depocenter, but the redistributed sedimentary
load produces displacement of the forebulge
toward the east relative to its early Paleocene
position (compare Figs. 14B and 14C). Thus
a broader late Paleocene basin is created. The
elevation of the depositional profile diminishes
from ~300 m in the western side of the basin to
0 m at the eastern side. Most of the upper Paleocene quiescent-phase strata are accommodated
between the undeformed reference datum and
the graded topographic profile in the western
side of the basin (west of 300 km). In the eastern part of the basin (east of 300 km), however,
most of these strata (e.g., Barco and Los Cuervos) are accommodated below the undeformed
reference datum in the space provided by flexural subsidence. This configuration is consistent
with a corresponding west-to-east change from
continental to coastal-plain facies.
In our fourth experiment, we chose the late
Eoceneearly Oligocene basinal configuration
in Figure 11A to test whether Eastern Cordillera
tectonic and sedimentary loading explain subsidence in the absence of Central Cordillera loading. This basin configuration postdated the final
episode of uplift and eastward expansion of the
Central Cordillera during the early Eocene. For
lack of constraints, we disregard the paleoelevation of the Central Cordillera slope in the western margin of the late Eoceneearly Oligocene
basin and assume a flat initial reference datum

TABLE 1. MECHANICAL PARAMETERS FOR 2-D FLEXURAL MODELING

t
i
i
m
g

70 Gpa
0.25
35 km
2400 kg/m3 (sedimentary rocks at shallow burials)
2700 kg/m3 (crystalline rocks, consolidated sedimentary rocks)
3300 kg/m3
9.8 m/s2

Note: EYoungs modulus, GpaGigapascals, Poissons ratio, t


elastic thickness, icrust density, mmantle density, gEarths gravity.

Figure 13. (A) Maastrichtian, (B) early Paleocene, and (C) late Paleocene discrete load configurations [h(x)] derived from Figure 9A. (D) Early Oligocene load configuration derived
from Figure 11A. See text for discussion.

Geological Society of America Bulletin, September/October 2005

1285

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

that corresponds to the surface of the Late Cretaceousearly Eocene foreland basin. The loads of
Eastern Cordillera sedimentary fill and folds in
Figure 13D produce the deformed configuration
in Figure 14D. The final configuration of the
early Oligocene deflection profile in the westernmost part of the basin (100160 km in horizontal scale of Fig. 14E) includes an assumed
paleoelevation of the Central Cordillera, which
restores the westward onlapping configuration
of strata. Thicknesses up to 400 m are accommodated between the undeformed reference
datum and the graded depositional profile in
the western part of the basin. The gradient of
the depositional profile decreases toward the
east, while flexural subsidence accommodates
all the sediment in the easternmost sector of the
basin (east of 500 km, Fig. 14E). In general, the
combination of flexural subsidence and lower
paleoaltitudes explains the persistent eustatic
signature of upper Eocenelower Miocene
strata in the Llanos Basin area (e.g., Figs. 14
and DR3, see footnote 1).
INTEGRATED INTERPRETATION
OF BASIN DEVELOPMENT AND
EXHUMATION OF THE CENTRAL
AND EASTERN CORDILLERAS
The successful fit between the results of the
two-dimensional mechanical models and the
patterns of sedimentary thicknesses and facies
east of the Central Cordillera indicate that the
first-order characteristics of the sedimentary fill
were simply controlled by flexural variations.
We can now extrapolate these results and assess
the degree to which regional distributions of
unconformities and facies were also determined
by similar mechanisms or by secondary controls
such as eustatic variation. The pictures of evolution in Figures 1517, DR4, and DR5 (see
footnote 1) illustrate this discussion.
Late CretaceousEarly Eocene Foreland
Basin

Figure 14. Modeled relative topography profiles [rt(x)] for (A) Maastrichtian, (B) early
Paleocene, (C) late Paleocene, and (D) and (E) early Oligocene times. These simulations satisfactorily replicate the subsidence mechanisms and basinal configurations first interpreted
from geological data sets. See text for discussion.

1286

Expanding Central Cordillera uplift created


a wedge-shaped foreland basin to the east with
regional drainage oriented toward the NE to the
Maracaibo Basin (Figs. 15 and 16; Campbell,
1968). Central Cordillera deformation was
episodic, which produced shifting positions of
flexural subsidence and marine-influenced sedimentation. At all times, the basin topographic
axis lay east of the maximum accommodation
areas, and it migrated toward the Central Cordillera during times of enhanced deformation
(Figs. 15 and 16, cross sections). Two periods
of uplift during the MaastrichtianDanian and
the early Eocene enhanced subsidence next to

Geological Society of America Bulletin, September/October 2005

Geological Society of America Bulletin, September/October 2005

FS: Fusagasug syncline


US: Usme syncline

Figure 15. Synthesis of (A) Maastrichtian and (B) early Paleocene (Danian) depositional environments, paleogeography, and flexural subsidence (in insets). See text for discussion. Main sources of data: Julivert (1963), Campbell (1968), Fabre (1983a, 1983b), Daz (1994), Sarmiento (1994), Cooper et al. (1995), Gmez et al. (2003, 2005).

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

1287

1288

Geological Society of America Bulletin, September/October 2005

FS: Fusagasug syncline


US: Usme syncline

Figure 16. Synthesis of (A) late Paleocene and (B) early Eocene depositional environments, paleogeography, and flexural subsidence (in insets). See text for discussion. Main
sources of data: Julivert (1963), Hoorn (1988), Cspedes and Pea (1995), Cooper et al. (1995), Cazier et al. (1995), and Gmez et al. (2003, 2005).

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

Geological Society of America Bulletin, September/October 2005

PA: Provincia anticline


LA: Lisama anticline
LC: Los Cobardes anticline
AA: Arcabuco anticline
VA: Villeta anticlinorium
GS: Guaduas syncline
FS: Fusagasug syncline
US: Usme syncline

Figure 17. Synthesis of (A) middle Eocene and (B) late Oligoceneearly Miocene depositional environments and paleogeography. See text for discussion and Figures DR4 and
DR5 (see text footnote 1) for the late Eoceneearly Oligocene and late Miocene reconstructions. Main sources of data: Brgl (1955), Hubach (1957), Julivert (1963), Restrepo
et al. (1975), Kellogg (1984), Ulloa (1985), Cooper et al. (1995), and Gmez et al. (2003, 2005).

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

1289

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.

the Central Cordillera, while erosive unconformities developed in the eastern sector of the
basin (Figs. 15A, 15B, and 16B). MaastrichtianDanian sea withdrawal was caused by increased
sediment supply from the Central Cordillera
and from exposed areas of the Guyana Shield
(Figs. 15A and 15B). Late Paleocene redistribution of sedimentary loads during Central
Cordillera tectonic quiescence created flexural
space for renewed coastal sedimentation in the
Llanos area (Fig. 16A), which filled an irregular
topography incised during the previous erosional period (e.g., Cazier et al., 1995). Ample
tectonic subsidence generated a narrower basin
in the early Eocene and facilitated the impact of
a sea-level rise (Haq et al., 1988) to be felt to the
north (upper Socha, Fig. 16B).
The Late CretaceousDanian unconformity
of the eastern sector of the foreland basin was
attributed previously to reverse faulting in the
Llanos foothills (Sarmiento, 1994; Villamil,
1999). However, this erosional surface has a
large geographic extent and cuts older strata
monotonically toward the Guyana Shield. Such
characteristics are diagnostic of regional-scale
controls of the kind exerted by regional flexure
rather than local faulting. The stratigraphic hiatuses associated with the MMVU of the MMVB
and the Llanos Paleogene unconformities
decrease toward the Eastern Cordillera, where
there are age-equivalent strata, albeit the strata
contain minor temporal gaps related to local
folding. Previous models correlated the MMVU
with the younger Paleogene unconformity of
the Llanos Basin across the Eastern Cordillera
and postulated a period of regional uplift and
erosion of all the Colombian territory during the
middle Eocene (Cooper et al., 1995; Pindell et
al., 1998; Villamil, 1999). However, the magnitude and duration of such an event are not consistent with preserved Maastrichtian-Oligocene
growth strata of the Eastern Cordillera, whose
formation required continuous sedimentation.
The sweeping regional-uplift interpretation also
makes it very difficult to explain the mechanical
causes of subsequent subsidence of such a vast
crustal uplift, which accommodated km-scale
thick piles of younger Cenozoic strata throughout the Colombian territory.
Middle Eocene to Neogene: Effects of
Eastern Cordillera Diachronous Uplift
No foreland basin coupled to the Central Cordillera has existed since middle Eocene times;
long-term subsidence in the MMVB and Llanos
Basin resulted from Eastern Cordillera sedimentary and tectonic loading. Crustal loading by
Eastern Cordillera anticlines and westward and
eastward onlapping sediments caused long-term,

1290

regional subsidence. The middle Eocene to early


Oligocene reconstructions (Figs. 17A and DR4,
see footnote 1) highlight the coexistence of two
different sedimentary systems, each with NEdirected drainage. Sedimentation in the west
(MMVB) was continental, a likely result of a
higher paleoaltitude and larger supply of sediment from the Central Cordillera. Farther to the
east, a marine transgression is recorded by facies
of the Eastern Cordillera and the Llanos Basin
(Cooper et al., 1995). The spatial parallelism of
these drainages is explained by an intervening
zone of deformation along the western half of
the Eastern Cordillera. The southern part of this
drainage divide was the Villeta anticlinorium, as
revealed by flanking middle EoceneOligocene
growth strata (Gmez et al., 2003).
During the late Eocene, shallow marine conditions were established to the east of the Eastern
Cordillera divide (Fig. DR4, see footnote 1). The
peak of this transgression broadly correlates with
a global sea-level highstand at the Eocene-Oligocene boundary (ca. 3334 Ma, Haq et al., 1988).
Its effect propagated into the northern MMVB
(Los Corros fossil horizon) from the Maracaibo
basin around the northern part of the Eastern
Cordillera divide (Fig. DR4, see footnote 1).
No physical evidence of this transgression exists
in the southernmost MMVB because this area
was at a higher altitude, and the growing Villeta
anticlinorium formed a barrier. Oligoceneearly
middle Miocene sedimentation of the Llanos
areas expanded toward the east (Fig. 17B) and
remained close to sea level (Carbonera Formation) due to the combination of low paleoaltitudes and creation of flexural accommodation by
tectonic and sedimentary loading.
Major changes in basin configuration happened during the late Oligoceneearly middle
Miocene due to deformation to the NE of the
MMVB (Gmez et al., 2005; Fig. 17B herein).
In palinspastically restored position, the Los
Cobardes, Provincia, and Lisama anticlines are
part of a larger morphostructural unit whose
northeastward continuation was the Perij
Range. Simultaneous deformation of the Perij
Range is documented by structural studies (Kellogg, 1984). We interpret that the Eastern CordilleraPerij RangeSantander Massif structural barrier further raised the MMVB base level
and forced the MMVB rivers to flow toward the
Llanos Basin region across the Eastern Cordillera region (Gmez et al., 2005). In the southern MMVB, lower to lower middle Miocene
sediments overlapped the Villeta anticlinorium
(Gmez et al., 2003, Fig. 17A herein). Another
global eustatic sea-level rise (Haq et al., 1988)
probably contributed to the early to early middle
Miocene tectonically enhanced accommodation
of the Colombian basins. The ensuing highstand

is recorded by the Len Formation (Llanos


Basin) and the slightly brackish deposits of the
Santa Teresa Formation (southern MMVB) and
the La Cira fossil horizon (northern MMVB).
The MMVB paleodrainage returned to the
north during the late middle to late Miocene
(Fig. DR5, see footnote 1) due to Eastern Cordillera continued uplift and sedimentary overlap
of the Cchira Arch (older northern boundary
of the MMVB, Gmez et al., 2005). Neogene
pulses of Eastern Cordillera uplift also caused
episodes of MMVB flexural tilting and partial
erosion of Neogene deposits (e.g., Fig. 2C).
CONCLUSIONS AND DISCUSSION
The Colombian foreland basin system east of
the Central Cordillera overlapped a Mesozoic
rift province, and subsequent inversion of the
rift system in the Eastern Cordillera modified
the nature of the foreland basin. Long-term
tectonic accommodation was controlled by
isostatic adjustments to variable distributions
of crustal tectonic loads, in combination with
sediment supply from tectonic highlands, and
paleoaltitude of depositional profiles. Eustasy
was a secondary factor, and its geographic
distribution was controlled by flexural accommodation and local deformation. Although
mechanically related, at least four different
types of unconformities can be recognized in
the foreland basin system, which reflect a range
of scales of variation in the wavelength of tectonic accommodation. First, the MMVU in the
western sector of the foreland system resulted
from eastward-migrating Late Cretaceousearly
Eocene Central Cordillera uplift and consequent
long-term erosion since the middle Eocene. Second, Paleogene unconformities formed on the
distal eastern side of the Late Cretaceousearly
Eocene foreland basin (e.g., Llanos area) due to
lithospheric flexure and erosional beveling during periods of Central Cordillera uplift. Third,
similar unconformities, but formed by eastward
tilting under the load of the Eastern Cordillera,
are found in the Neogene MMVB. Fourth,
shorter wavelength folding produced local
growth unconformities and modified drainage
patterns in the subsiding basin between the
MMVB and Llanos area. Growth strata are
the most important evidence of Late CretaceousNeogene diachronous Eastern Cordillera
deformation, prior to massive Pliocene uplift.
The two-dimensional mechanical models
seem to explain variations in accommodation
at a broad regional level in the palinspastically
restored area east of the Central Cordillera,
which suggests that mechanical controls were
relatively simple during the Cenozoic. Thus,
there is no need to utilize alternative modeling

Geological Society of America Bulletin, September/October 2005

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


DEVELOPMENT OF THE COLOMBIAN FORELAND-BASIN SYSTEM AND NORTHERN ANDES

techniques that invoke broken plates or thermal


weakening of the lithosphere. These models are
better suited for cases in which simple flexural
approaches cannot explain complex distributions
of sedimentary thicknesses and facies, as exemplified by foreland basins flanking the Central
Andes (e.g., Cardozo and Jordan, 2001).
This paper highlights the importance of
assessing the mechanical causes of accommodation and unconformities to understand the
evolution of tectonic basins. Time correlations
based on sequence stratigraphy, if they assume
simultaneous base-level variations that are sensitive to choice of datum, are not appropriate for
basin analysis at scales comparable to the distances over which tectonic controls act. For the
MMVBEastern CordilleraLlanos system, we
find that there are four different scales of tectonic
control on unconformities. At each of these spatial scales, in relatively distal basin areas far from
zones of active deformation, it may be difficult to
recognize the tectonic control on unconformities
and correlative conformities. It would be just as
easy to ascribe these surfaces to eustasy even if
that explanation were erroneous. This discussion
illustrates the importance of placing the detailed
local studies within a thorough regional study, at
which scale the tectonic controls are clear.
ACKNOWLEDGMENTS

This study was supported by grants and fellowships from the Petroleum Research Fund (ACS-PRF
no. 32818-AC8), the National Science Foundation
(Faculty Award to Women in Science and Engineering
award GER-9022811 to T.E. Jordan), and the Instituto
Colombiano para el desarrollo de la Ciencia y la Tecnologa (Colciencias). Cornell University, the Geological Society of America, the American Association of
Petroleum Geologists, Shell Oil Company Foundation,
Shell E&P Solutions, and Ecopetrol also contributed
funding to this research. We also thank geologists
Matthew Burns and Francisco Gmez for helpful
discussions. Critical reviews by Rebecca Dorsey,
Brian Horton, Ken Ridgway, Allen Glazner, Cynthia
Evinger, Frdric Mouthereau, Yildirim Dilek, and an
anonymous reviewer helped us to improve this paper.
REFERENCES CITED
Acosta, J.E., and Beltrn, W.E., 1987, Estratigrafa de la
Formacin La Regadera en el flanco Occidental del
Sinclinal de Usme [B.S. thesis]: Bogot, Universidad
Nacional de Colombia, 65 p.
Allen, P., and Allen, J., 1992, Basin Analysis, Principles and
Applications: London, Blackwell Scientific Publications, 451 p.
Anadn, P., Cabrera, L., Ferrn, C., Marzo, M., and Riba, O.,
1986, Syntectonic intraformational unconformities in
alluvial fan deposits, eastern Ebro Basin margins (NE
Spain), in Allen, P., and Homewood, P., eds., Foreland
basins: International Association of Sedimentologists
Special Publication 8, p. 259271.
Anders, M.H., Krueger, S.W., and Sadler, P.M., 1987, A new
look at sedimentation rates and the completeness of the
stratigraphic record: Journal of Geology, v. 95, p. 114.
Aspden, J.A., McCourt, W.J., and Brook, M., 1987, Geometrical control of subduction-related magmatism:

The Mesozoic and Cenozoic plutonic history of


western Colombia: Journal of the Geological Society,
v. 144, p. 893905.
Blair, T.C., and McPherson, J.G., 1994, Alluvial fans and
their natural distinction from rivers based on morphology, hydraulic processes, sedimentary processes, and
facies assemblages: Journal of Sedimentary Research,
v. A64, p. 450489.
Brgl, H., 1955, Globorotalia fohsi en la Formacin de
Usme: Boletn Geolgico Servicio Geolgico Nacional, no. 3, p. 5665.
Campbell, C.J., 1968, The Santa Marta wrench fault of
Colombia and its regional setting, in Fourth Caribbean
Geological Conference, 1965: Trinidad, Port of Spain,
p. 247261.
Cardozo, N., and Jordan, T., 2001, Causes of spatially
variable tectonic subsidence in the Miocene Bermejo
foreland basin, Argentina: Journal of Basin Research,
v. 13, p. 335357.
Cazier, E.C., Hayward, A.B., Espinosa, G., Velandia, J.,
Mugniot, J.F., and Leel, W.G., 1995, Petroleum Geology of the Cusiana Field, Llanos Basin Foothills,
Colombia: American Association of Petroleum Geologists Bulletin, v. 79, p. 14441463.
Cspedes, S., and Pea, L., 1995, Relaciones estratigrficas y ambientes de depsito de las formaciones del
Terciario Inferior aflorante entre Tunja y Paz de Ro
(Boyac) [B.S. thesis]: Bogot, Universidad Nacional
de Colombia, 74 p.
Colletta, B., Hebrard, F., Letouzey, J., Werner, P., and
Rudkiewicz, J.L., 1990, Tectonic style and crustal structure of the Eastern Cordillera (Colombia) from a balanced
cross section, in Letouzey, J., ed., Petroleum and Tectonics in Mobile Belts: Paris, Editions Technip, p. 81100.
Cooper, M.A., Addison, F.T., Alvarez, R., Coral, M., Graham, R.H., Hayward, A.B., Howe, S., Martnez, J.,
Naar, J., Peas, R., Pulham, A.J., and Taborda, A.,
1995, Basin development and tectonic history of the
Llanos Basin, Eastern Cordillera and Middle Magdalena Valley, Colombia: American Association of Petroleum Geologists Bulletin, v. 79, p. 14211443.
Cortina, J.D., and Valvuena, J.A., 1987, Anlisis paleoambiental de la Formacin Arenisca del Cacho en el sinclinal Checua-Lenguazaque, Cundinamarca [B.S. thesis]:
Bogot, Universidad Nacional de Colombia, 85 p.
Crowley, K.D., 1984, Filtering of depositional events and
the completeness of sedimentary sequences: Journal of
Sedimentary Petrology, v. 54, p. 127136.
Cuervo, E.A., and Ramrez, A., 1985, Estratigrafa y ambiente de sedimentacin de la Formacin Cacho en los
alrededores de Bogot [B.S. thesis]: Bogot, Universidad Nacional de Colombia, 75 p.
Dengo, C.A., and Covey, M.C., 1993, Structure of the
Eastern Cordillera of Colombia: Implications for trap
styles and Regional Tectonics: American Association
of Petroleum Geologists Bulletin, v. 77, p. 13151337.
Daz, L., 1994, Reconstruccin de la cuenca del Valle Superior del Magdalena, a finales del Cretcico, in Etayo
Serna, F., ed., Estudios Geolgicos del Valle Superior
del Magdalena: Universidad Nacional de ColombiaEcopetrol Publicacin Especial, p. XI1XI13.
Etayo-Serna, F., 1964, Posicin de las faunas en los depsitos
Cretcicos Colombianos y su valor en la subdivisin
cronolgica de los mismos: Boletn de Geologa Universidad Industrial de Santander, no. 16-17, p. 5141.
Etayo-Serna, F., 1985, Paleontologa estratigrfica del
Sistema Cretcico en la Sierra Nevada del Cocuy,
in Etayo-Serna, F., and Laverde, F., eds., Proyecto
Cretcico: Publicaciones Geolgicas Especiales del
Ingeominas, no. 16, p. XXIV1XXIV47.
Etayo-Serna, F., Barrero, D., Lozano, H., Espinosa, A.,
Gonzlez, H., Orrego, A., Ballesteros, I., Forero, H.,
Ramrez, C., Zambrano, F., Duque, H., Vargas, R.,
Nez, A., Alvarez, J., Ropan, C., Cardozo, E.,
Galvis, N., Sarmiento, L., Albers, J., Case, J.,
Singer, D., Bowen, R., Berger, B., Cox, D., and
Hodges, C., 1983, Mapa de Terrenos Geolgicos de
Colombia: Publicaciones Geolgicas Especiales del
Ingeominas, no. 14, 235 p.
Fabre, A., 1983a, La subsidencia de la Cuenca del Cocuy
(Cordillera Oriental de Colombia) durante el Cretceo
y el Terciario Segunda Parte: Esquema de evolucin
tectnica: Geologa Norandina, no. 8, p. 2227.

Fabre, A., 1983b, La subsidencia de la Cuenca del Cocuy


(Cordillera Oriental de Colombia) durante el Cretceo
y el Terciario primera parte estudio cuantitativo de la
subsidencia: Geologa Norandina, no. 8, p. 4961.
Fabre, A., 1985, Dinmica de la sedimentacin Cretcica
en la regin de la Sierra Nevada del Cocuy (Cordillera
Oriental de Colombia), in Etayo-Serna, F., and Laverde,
F., eds., Proyecto Cretcico: Publicaciones Geolgicas
Especiales del Ingeominas, no. 16, p. XIX1XIX20.
Fabre, A., and Delaloye, M., 1982, Intrusiones bsicas Cretcicas de la Cordillera Oriental: Geologa Norandina,
no. 6, p. 1928.
Flemings, P.B., and Jordan, T.E., 1990, Stratigraphic modeling of foreland basins: Interpreting thrust deformation
and lithospheric rheology: Geology, v. 18, p. 430435.
Ford, M., Williams, E.A., Artoni, A., Verges, J., and
Hardy, S., 1997, Progressive evolution of a faultrelated fold pair from growth strata geometries, Sant
Lloren de Morunys, SE Pyrenees: Journal of Structural Geology, v. 19, p. 413441.
Geotec Ltda, 1988, Geologic Map of Colombia, second edition: Geotec Ltda, scale 1:1,200,000, 1 sheet.
Germeraad, J.H., Hopping, C.A., and Muller, J., 1968, Palynology of Tertiary sediments from tropical areas: Reviews of
Palaeobotany and Palynology, v. 6, p. 189348.
Gmez, E., 2001, Tectonic controls on the Late Cretaceous
to Cenozoic sedimentary fill of the Middle Magdalena
Valley Basin, Eastern Cordillera and Llanos Basin,
Colombia [Ph.D. thesis]: Ithaca, New York, Cornell
University, 619 p.
Gmez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K.,
Kelley, S., and Heizler, M., 2003, Controls on architecture of the Late Cretaceous to Cenozoic Southern Middle Magdalena Valley Basin, Colombia: Geological
Society of America Bulletin, v. 115, no. 2, p. 131147.
Gmez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K.,
and Kelley, S., 2005, Syntectonic Cenozoic sedimentation in the Northern Middle Magdalena Valley Basin
and implications for exhumation of the northern
Andes: Geological Society of America Bulletin, v. 117,
p. 547569.
Haq, B.U., Hardenbol, J., and Vail, P.R., 1988, Mesozoic
and Cenozoic chronostratigraphy and cycles of sealevel change, in Wilgus, C.K., et al., eds., Sea-level
changes: An integrated approach: Society of Economic
Paleontologists and Mineralogists Special Publication 42, p. 71108.
Helmens, K.F., 1990, Neogene-Quaternary geology of the
high plain of Bogot, Eastern Cordillera, Colombia
(Stratigraphy, paleoenvironments and landscape evolution), in Dissertationes Botanicae, Band 163: Stuttgart,
J. Cramer Verlag, 202 p.
Hoorn, C., 1988, Quebrada del Mochuelo, type locality
of the Bogot Formation, a sedimentological, petrographical, and palynological study: Hugo de Vries
Laboratory, University of Amsterdam, 21 p.
Hubach, H., 1957, Estratigrafa de la Sabana de Bogot y
alrededores: Boletn Geolgico Instituto Geolgico
Nacional, v. 5, no. 2, p. 93112.
Jaramillo, L., Roa, E., and Torres, M., 1993, Relaciones estratigrficas entre las unidades arenosas del
Palegeno (Paleoceno) del piedemonte llanero y la
parte media de la Cordillera Oriental [B.S. thesis]:
Bogot, Universidad Nacional de Colombia, 44 p.
Jordan, T.E., 1981, Thrust loads and foreland basin evolution, Cretaceous, western United States: American
Association of Petroleum Geologists Bulletin, v. 65,
p. 25062520.
Jordan, T.E., 1995, Retroarc foreland and related basins,
in Busby, C.J., and Ingersoll, R.V., eds., Tectonics of
sedimentary basins: Malden, Massachusetts, Blackwell
Science, p. 331362.
Julivert, M., 1958, La morfoestructura de la zona de mesas
al SW de Bucaramanga (Colombia S.A.): Boletn de
Geologa Universidad Industrial de Santander, no. 1,
p. 943.
Julivert, M., 1963, Los rasgos tectnicos de la regin de
la Sabana de Bogot y los mecanismos de formacin
de las estructuras: Boletn de Geologa Universidad
Industrial de Santander, no. 13-14, p. 5102.
Kellogg, J.N., 1984, Cenozoic tectonic history of the Sierra
de Perij, Venezuela-Colombia, and adjacent basins, in
Bonini, W.E., Hargraves, R.B., and Shagam, R., eds.,

Geological Society of America Bulletin, September/October 2005

1291

Downloaded from gsabulletin.gsapubs.org on December 15, 2012


GMEZ et al.
The CaribbeanSouth America Plate Boundary and
Regional Tectonics: Geological Society of America
Memoir 162, p. 239261.
Kellogg, J.N., Salvador, M., and Ojeda, G., 1995, Lithospheric structure of the Colombian Andes, flexure and
crustal shortening: Eos (Transactions, American Geophysical Union), v. 76, no. 46, Supplement, p. 374.
Laverde, F., 1989, Stratigraphy of the Tertiary sequence
Southwest of Bogot, Colombia, Northeastern Upper
Magdalena Valley, Western border of the Cordillera
Oriental [M.S. thesis]: Chapel Hill, University of
South Carolina, 66 p.
Macellari, C., 1988, Cretaceous paleogeography and depositional cycles of western South America: Journal of
South American Earth Sciences, v. 1, p. 373418.
Pindell, J.L., 1993, Regional synopsis of Gulf of Mexico and
Caribbean evolution, in Pindell, J.L., and Perkins, B.F.,
eds., 13th Annual Research Conference Proceedings,
Mesozoic and Early Cenozoic development of the Gulf
of Mexico and Caribbean region: Gulf Coast Section
Society of Economic Paleontologists and Mineralogists Foundation, p. 251274.
Pindell, J.L., Higgs, R., and Dewey, J.F., 1998, Cenozoic
palinspastic reconstruction, paleogeographic evolution, and hydrocarbon setting of the northern margin of
South America in Pindell, J.L., and Drake, C.L., eds.,
Paleogeographic evolution and non-glacial eustasy,
northern South America: Society for Sedimentary
Geology (SEPM) Special Publication 58, p. 4585.
Raasveldt, H.C., 1956, Mapa geolgico de la Repblica de
Colombia, Plancha L-9 (Girardot): Instituto Geolgico
Nacional, scale 1:200,000, 1 sheet.
Restrepo, G., Cepeda, H., and Muoz, A., 1975, Contribucin al conocimiento de los yacimientos del hierro

1292

de Paz de Ro (petrognesis y facies) [B.S. thesis]:


Bogot, Universidad Nacional de Colombia, 68 p.
Rodrguez, E., and Ulloa, C., 1994a, Geologa de la Plancha 169-Puerto Boyac: Bogot, Minera y Qumica,
Memoria resumida, Instituto de Investigaciones en
Geociencias, 31 p.
Rodrguez, E., and Ulloa, C., 1994b, Geologa de la Plancha
189-La Palma: Bogot, Minera y Qumica, Memoria
resumida, Instituto de Investigaciones en Geociencias,
57 p.
Roeder, D., and Chamberlain, R.L., 1995, Eastern Cordillera of Colombia: Jurassic-Neogene Crustal Evolution, in Tankard, A.J., Surez, R., and Welsink, H.J.,
eds., Petroleum basins of South America: American
Association of Petroleum Geologists Memoir 62,
p. 633645.
Rowan, M.G., and Linares, R., 2000, Fold-evolution matrices and axial-surface analysis of fault-bend folds:
Application to the Medina anticline, Eastern Cordillera, Colombia: American Association of Petroleum
Geologists Bulletin, v. 84, p. 741764.
Sarmiento, G., 1994, Estratigrafa, palinologa, y paleoecologa
de la Formacin Guaduas: Publicaciones Geolgicas
Especiales del Ingeominas, no. 20, p. 1192.
Sarmiento, L.F., 1989, Stratigraphy of the Cordillera Oriental, West of Bogot, Colombia [M.S. thesis]: Chapel
Hill, University of South Carolina, 102 p.
Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A.,
Herv, P., Harmen, B., Olaya, J., and Rivera, C., 2000,
Geodynamics of the northern Andes: Subductions and
intracontinental deformation (Colombia): Tectonics,
v. 19, p. 787813.
Toro, J., 1991, The termination of the Bucaramanga fault
in the Cordillera Oriental, Colombia [abstract]: Vth

International Circumpacific terrane conference, Serie


Comunicaciones, Departamento de Geologa, Facultad
de Ciencias Fsicas y Matemticas, Universidad de
Chile, no. 42, p. 226.
Turcotte, D.L., and Schubert, G., 1982, Geodynamics,
Applications of Continuum Physics to Geological
Problems: New York, John Wiley and Sons, 450 p.
Ulloa, C.E., 1985, Hierro ooltico en el norte de Sur
Amrica: Transactions of the Fourth Latin American
Geological Conference, v. 4, p. 263275.
Van der Hammen, T., 1957, Estratigrafa palinolgica de la
Sabana de Bogot (Cordillera Oriental de Colombia):
Boletn Geolgico del Servicio Geolgico Nacional,
v. 5, no. 2, p. 189203.
Vergara, L.E., and Rodrguez, G., 1997, The Upper Cretaceous and Lower Paleocene of the Eastern Bogot
Plateau and Llanos thrust belt, Colombia: Alternative
appraisal to the nomenclature and sequence stratigraphy: Geologa Colombiana, no. 22, p. 5179.
Villamil, T., 1999, Campanian-Miocene tectonostratigraphy, depocenter evolution and basin development of
Colombia and western Venezuela: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 153, p. 239275.
Villegas, M.E., Bachu, S., Ramn, J.C., and Underschultz,
J.R., 1994, Flow of formation waters in the CretaceousMiocene succession of the Llanos Basin, Colombia:
American Association of Petroleum Geologists Bulletin, v. 78, p. 18431862.
MANUSCRIPT RECEIVED BY THE SOCIETY 22 JUNE 2003
REVISED MANUSCRIPT RECEIVED 18 OCTOBER 2004
MANUSCRIPT ACCEPTED 21 OCTOBER 2004
Printed in the USA

Geological Society of America Bulletin, September/October 2005

S-ar putea să vă placă și