Vous êtes sur la page 1sur 69

Structural Geometry of the Jura-Cretaceous Rift of the Middle

Magdalena Valley Basin Colombia






Luisa F. Rolon



Thesis submitted to the Eberly College of Arts and Sciences
at West Virginia University
in partial fulfillment of the requirements for the degree of



Master of Science in Geology




Committee:
Jaime Toro, Ph.D., Chair
Thomas Wilson, Ph.D.


Department of Geology and Geography




Morgantown, WV
2004





keywords: rift basins, normal fault, inversion, half graben, Middle Magdalena
Valley basin, seismic interpretation.
ABSTRACT

Structural Geometry of the Jura-Cretaceous Rift of the Middle Magdalena
Valley Basin Colombia

Luisa F. Rolon


The Middle Magdalena Valley (MMV) basin, Colombia, evolved to its actual
configuration through several stages closely linked with the tectonic events of the
northwest corner of South America.
Early J urassic extensional deformation related to the break-up of Pangea, resulted
in the development of rift structures in the MMV basin area. The syn-rift infill consists of
fluvial and lacustrine sedimentary rocks (J ordn, Girn and Santos formations), overlain
by limestones and shales (Basal Lime Group). Irregular discontinuous reflectors, rotated
by normal faults and increasing thickness towards the south-east, characterize the J urassic
syn-rift section. In Early Aptian, a post-rift phase controlled by thermal subsidence
started. During this phase the Simit, La Luna and Umir formations were deposited. The
seismic character of this sequence consists of continuous reflections and gradual changes
in thickness. From the end of the Cretaceous through the present, the collision of oceanic
terranes against the western margin of Colombia caused thrusting in the Central and
Eastern Cordilleras. The MMV basin became an intermontane basin. Sedimentation
during this period consisted of fluvial deposits accumulated as continental molasses.
I propose an asymmetrical half graben geometry for the J ura-Cretaceous structure
of the northern part of the MMV basin. The range-bounding faults of the Eastern
Cordillera: the Bucaramanga, La Salina, and Cambao-Bituima faults were originally
normal faults that controlled thickness of the synrift sequence and, in general, the
geometry of the structure.
The rift structures are well preserved in the study area. However, most of the
preserved faults were inverted during the Tertiary orogeny. Inversion can be documented
by structures such as shortcuts, thicker rift strata on the hanging wall, and by variation of
fault throw.

Acknowledgments

I give thanks to God for giving me the means, the skills and the courage to complete my
thesis and my degree, and to my parents for their continuous and unconditional support.

I also would like to express my appreciation to Dr. Thomas Wilson for his valuable
comments and suggestions and in particular I would like to extend my deepest gratitude
to my research advisor and committee chairman, Dr. J aime Toro, for his continuous
guidance through this work. Gracias Doc.

I want to express my gratitude to Ecopetrol and La Luna Oil Company for providing
valuable information used for this work and in the same way I want to send special
thanks to Dr. Dario Barrero for his contributions and comments.

Thanks a lot to the Department of Geology and Geography of West Virginia University
for giving me the opportunity and the means to complete my Masters degree.

I want to thanks to my boyfriend Bret McDaniel for giving me moral and constant
support during the most difficult moments of this project.

And last but not the least I would like to thank all my friends at West Virginia University
who one way or another helped me in my work.
iii
TABLE OF CONTENTS
Chapter Page

1. INTRODUCTION.......................................................................................................... 1
2. GEOLOGICAL SETTING............................................................................................. 3
2.1. Structural Geology................................................................................................... 3
2.2. Stratigraphy.............................................................................................................. 4
2.3. Tectonic Evolution................................................................................................... 5
2.3.1. Upper J urassic Campanian-Maastrichtian: Terrestrial Rift........................... 5
2.3.1.2. Middle Albian - Maastrichtian: Post-rift phase (thermal subsidence)........... 7
2.3.2. Maastrichtian - Oligocene: Retroarc Foreland Basin........................................ 8
2.3.3. Miocene - Recent: Broken Foreland (Intermontane) Basin development........ 9
3. BACKGROUND ON RIFT BASINS........................................................................... 10
3.1. Structure................................................................................................................. 10
3.1.1. Initial Aperture................................................................................................ 12
3.1.2. Synrift Subsidence.......................................................................................... 13
3.1.3. Inversion of extensional structures................................................................. 14
3.2. Stratigraphy............................................................................................................ 15
3.2.1. Early synrift deposits...................................................................................... 16
3.2.2. Late synrift deposits........................................................................................ 17
3.2.3. Synrift unconformities.................................................................................... 18
4. INTERPRETATION/RESULTS.................................................................................. 19
4.1. Syn-rift stratigraphy............................................................................................... 21
5. DISCUSSION............................................................................................................... 24
6. CONCLUSIONS........................................................................................................... 27
REFERENCES CITED..................................................................................................... 29

iv

LIST OF FIGURES
Figure 1. Location of the study area.
Figure 2. Distribution of wells and 2D seismic data used.
Figure 3. Geologic map of the Middle Magdalena Valley.
Figure 4. Cross sections through the northern Middle Magdalena Valley Basin.
Figure 5a. Generalized stratigraphic column.
Figure 5b Subsurface expression of units studied in this work.
Figure 6. Isopach maps in time for acoustic units in the East African Rift System.
Figure 7. Diagram of asymmetric half-graben profiles.
Figure 8. Diagram of different geometries of asymmetric half-graben structures.
Figure 9. Accommodation zones of the Gulf of Suez.
Figure 10. Schematic diagram of a positive inversion structure.
Figure 11. Model for development of a shortcut thrusting.
Figure 12. Interpretation of the seismic line BSM91-2270.
Figure 13. Interpretation of the seismic line B89-1370.
Figure 14. Interpretation of the seismic line BSM91-2020E.
Figure 15. Interpretation of the seismic line BSM91-1420.
Figure 16. Interpretation of the seismic line NC-81-2.
Figure 17. Structural map in time at the base of the J urassic wedge.
Figure 18. Structural map in time at the top of the J urassic wedge.
Figure 19. Structural map in time at the base of the Basal Lime Group.
Figure 20. Electrofacies of the Arcabuco-Los Santos formations in the study area (see
figure 2 for location of wells).
Figure 21. Isopach map (in time) of the J urassic wedge.
Figure 22. Isopach map (in time) of the Basal Lime Group.
Figure 23. Stratigraphic cross-section along the study area.
Figure 24. Valanginian Paleo-geographic model of the Magdalena Valley and the
Eastern Cordillera.
v
Figure 25. Comparison of the geometry of the rift of the MMVB with model
proposed by Rosendahl et al. (1986).
Figure 26. Model of the J urassic structural geometry of the Middle Magdalena Valley
Basin proposed in this thesis.
vi
1. INTRODUCTION
The Colombian Andes consist of three separate ranges: (1) the Western Cordillera,
composed of accreted oceanic rocks, (2) the Central Cordillera composed of Phanerozoic
igneous and metamorphic rocks including an active volcanic arc, and (3) the Eastern
Cordillera, most of which is underlain by uplifted J urassic through Tertiary sedimentary
rocks and their Precambrian through Mesozoic basement. The Romeral Suture Zone,
which extends from the Guayaquil Gulf in Ecuador, to the Caribbean Sea in northern
Colombia, separates the accreted oceanic terranes of the Western Cordillera from
continental terranes of the Central and Eastern cordilleras (Barrero et al., 1969; Barrero,
1979) (figure 1).
The Middle Magdalena Valley basin, located in the central part of Colombia
between the Central Cordillera to the west and the Eastern Cordillera to the east,
constitutes a poly-historic basin, that evolved to its present configuration through different
stages closely linked with the tectonic events of the northwest corner of South America.
These events range from a rift stage to tectonic inversion. These stages are related to the
interaction of Faralln/Nazca, South American and Caribbean tectonic plates. Thus, the
final result of this long history is a complex structural and stratigraphic framework.
During Triassic- Early Cretaceous time, as the Atlantic opened, the MMVB went through a
rifting phase. Although this event has been mentioned previously in several publications
about the evolution of the MMV basin and the Eastern Cordillera (e.g. Cooper et al., 1995;
Dengo and Covey, 1993; Morales et. al., 1958; Gomez, 2001), little documentation is
provided, thus the structural geometry and the sequence stratigraphy of the rift are poorly
1
known. The reasons for the lack of attention to the pre-mid Cretaceous history of the basin
include the abrupt topography in the foothills of the Eastern Cordillera, the almost total
absence of outcrops in the valley, and the lack of interest by the petroleum industry, the
main sponsor for geologic studies in the area. Thus far petroleum exploration has focused
on the mid-Cretaceous to Tertiary rocks. Nevertheless, there is great potential for new
hydrocarbon plays in the neglected synrift sequence. Understanding of the geometry and
kinematics of this rift, as well as its infill cycles, will provide new insight into the tectonic
evolution of the area and better understanding about the generation, migration, and
trapping of the hydrocarbons in the basin.
This work is aimed at establishing a geometrical model of the main features of the
Colombian J urassic-Cretaceous rift in the area occupied by present northern part of the
Middle Magdalena Valley. My research is based on approximately 300 kilometers of 2D
seismic lines, 42 well logs that reached the early Cretaceous rocks (figure 2), a surface
geology map scale 1:200,000, and 52 stratigraphic columns, provided by Ecopetrol and La
Luna Oil Co. Seismic data were used to identify the faults associated with the rift structure.
This thesis also addresses the sedimentology of the synrift and post-rift (thermal-
subsidence phase) sequences. Well logs, either on or close to the seismic lines, and surface
data were used to calibrate markers that represent sequence boundaries. As a result of these
interpretations structural and isopach maps were prepared.
I interpret the study area as a major half graben, striking NE-SW and tilted toward
the southeast. This hypothesis is based on the presence of some extensional structures
2
aligned in NE-SW direction and the southeastward thickening of the synrift sequence
(J urassic - Lower Cretaceous).
South of the study area, adjacent to the Bituima fault system, the J urassic sequence
thickens to the west (previously documented with seismic data in Cediel, Barrero, and
Caceres, 1998). This suggests that the vergence of the rift-bounding fault switches from
down-to-the-NW in the north to down-to- the-east in the southern MMV basin.

2. GEOLOGICAL SETTING
2.1. Structural Geology
The Middle Magdalena Valley basin is bounded to the west and south by two major
right-lateral transcurrent faults known as the Palestina and Ibague faults respectively. In
the northeast side, it is bounded by the Santander Massif and by the Bucaramanga fault, a
major Tertiary strike-slip fault, that trends northwest and ends in the Eastern Cordillera
thrust belt (Toro, 1990) (figure3).
In the study area, the west side of the basin is a monocline dipping toward the east,
constituted by a Paleozoic Precambrian basement that is overlain by the J urassic-
Cretaceous sedimentary wedge which is gradually onlapping toward the west (figures 4a
and 4b). The whole section is progressively eroded by a series of Tertiary unconformities,
that range in age from Paleocene to Miocene. In the southern part of the area, the basement
is cut by high-angle reverse faults, apparently related with the Palestina fault system, that
chiefly affected the western part of the basin during Late Paleocene (Ecopetrol, 2001)
(figure 4c).
3
The eastern boundary of the basin has a more complex structural style (figure 4)
consisting of a west-verging thrust belt and a swarm of high-to-low angle inverted normal
faults generally dipping toward the east. These faults are the main contributors to the uplift
of the Eastern Cordillera (Cooper et al., 1995). The thrusts are rooted in the Eastern
Cordillera basement and extend within the basin as far as the Magdalena River (figure 3).
Transport direction varies from WNW in the central sector to a more northwesterly
direction in the southern sector of the basin (Rolon et al., 2001). Thrust advance is uneven
and has not been continuous in time. The La Salina fault system constitutes the main
range-bounding fault in the study area.
2.2. Stratigraphy
The stratigraphic nomenclature of the Middle Magdalena Valley Basin that has
been widely used and is still valid, was first described by Morales et al. (1958) and
subsequently modified by Etayo and Laverde (1985). The stratigraphic column consists of
an igneous-metamorphic basement that possibly dates from Precambric to Lower
Paleozoic, overlain by a sequence of Late J urassic to recent sedimentary rocks. The main
characteristics of the sedimentary succession can be summarized as follows (figures 5a):
1) J urassic and Lower Cretaceous (Berriasian) volcano-sedimentary rocks
deposited in a tectonically controlled fluvial system, (Girn, and Arcabuco-Los Santos
formations) (Mojica et al., 1984; Etayo, 1968; Fabre, 1983). This units are overlain by
Early Cretaceous calcareous and siliciclastic sedimentary rocks deposited on a shallow
marine platform, that was wider than the present Middle Magdalena Valley (Basal Lime
4
Group conformed by the Cumbre, Rosablanca, Paja, and Tablazo formations) (figures 5a
and 5b).
2) Albian through Maastrichtian shallow to deep marine sedimentary rocks
accumulated on a platform affected by eustatic sea level changes (Simit, and La Luna
formations).
3) Upper Cretaceous to Paleocene deep marine siliciclastic to paludal rocks (Umir
and Lisama formations) deposited in a transgressive - regressive cycle during the first
stages of the Tertiary compressional deformation (figure 5a).
4) Tertiary fluvial and lacustrian clastic rocks (La Paz, Esmeraldas, Mugrosa,
Colorado and, Real formations), deposited during the progressive uplift of the Central and
Eastern cordilleras (Coletta et al., 1990; Dengo and Covey, 1993; Cooper et al., 1995)
(figure 5a).
2.3. Tectonic Evolution
2.3.1. Upper Jurassic Campanian-Maastrichtian: Terrestrial Rift
2.3.1.1. Upper J urassic Aptian: Syn-rift phase (extension)
In very late Triassic (Norian Retinian) to Early J urassic the opening of the
westernmost Tethys and the break-up of Pangea caused the initial separation of South
America, North America, and Africa. The extensional stresses within the continental block
resulted in the formation of a northwest - southeast trending rift structure in the MMV area
(Rolon et al., 2001; Etayo, Barrero and Renzoni, 1969; Etayo et al., 1983; Fabre, 1983).
According to Fabre (1983), the rifting process was passive, and it is described as an east-
west extension phase that caused crust and lithosphere to thin. Due to the elevated position
5
of the asthenosphere lithosphere boundary under thinned areas there was initial partial-
mantle-fusion and consequently, the emplacement of basic intrusions during J urassic
Cretaceous time. Heat flow during this phase was high (Fabre, 1983). Simultaneously, the
Upper Magdalena Valley, was scenario for important granitic and granodioritic plutonism
(Mojica and Dorado, 1987). Because of this magmatic belt many authors have interpreted
the later stage of the evolution of the MMV basin as a back-arc basin (Mojica and Dorado,
1987; Schamel, 1989; Dengo and Covey, 1993; Cooper et al., 1995). However, other
authors interpret the later stage of the extension phase as an aborted rift or an aulacogen
(Rolon et al., 2001; Etayo, Barrero and Renzoni, 1969; Etayo et al., 1983; Fabre, 1983). To
date there is not enough evidence to decide whether the most suitable model for this stage
of the basin evolution is an aulacogen or a back-arc basin.
Infill of the rift during the synrift stage represents a first tectono-sequence, which is
conformed by the J ordn, Girn, Santos, and the Basal Lime Group (Cumbre, Rosablanca,
Paja and Tablazo formations). Abrupt changes in thickness of these units suggest that their
accumulation took place while normal faults were still active.
Fluvial, lacustrine sedimentary rocks of the Girn, J ordn, and Santos formations
represent the lowstand of this sequence. In turn, limestones and silisiclastic shales of the
Cumbre, Rosablanca, Paja and Tablazo formations (Basal Lime Group) represent the
transgressive and the highstand systems tracts respectively (the transgressive systems tract
is constituted by the Cumbre and Rosablanca formations and the highstand systems tract is
represented by the Paja and Tablazo formations). At the top of this sequence a regional
6
unconformity is developed, which could be interpreted as the result of a first thermal decay
phase (Roeder and Chamberlain, 1995).
2.3.1.2. Middle Albian - Maastrichtian: Post-rift phase (thermal subsidence)
By the Middle Albian the extensional deformation had decreased and declining
heat flow lead to the beginning of a post-rift phase controlled by thermal subsidence. The
post-rift phase was characterized by the formation of a vast marginal sag; the dominant
deformation style was still normal faulting but less pervasive than during the synrift phase.
Accommodation space was created by thermal subsidence in addition to a global eustatic
rise that occurred during this period (Rolon and Carrero, 1995; Sarmiento, 2001; Villamil,
1999). During this phase, thickness of units became more uniform along the basin. Pelagic
material accumulated at relatively high rates due to upwelling-enhanced productivity
produced very good quality source rocks (Pindell and Tabbutt, 1995). The infill of the
basin is represented by three 3
rd
-order sequences that are very well documented (Villamil,
1999). Units included in this interval are: Simit, La Luna, and Umir formations. All these
sequences are type 2 constituted by a transgressive systems tract, followed by a highstand
systems tract. There are two events of maximum flooding, one during Middle Albian
represented by the Simit formation and a second one during Turonian-Coniacian time,
which is represented by La Luna formation. La Luna formation constitutes the best source
rock, not only in the MMV basin but also in other areas of Colombia and even in
Venezuela and Ecuador. The origin of the boundaries that frame the aforementioned
sequences is not clear yet, however, the most commonly accepted explanation is that they
represent eustatic changes of sea level (Villamil, 1999).
7
2.3.2. Maastrichtian - Oligocene: Retroarc Foreland Basin
At the end of the Cretaceous (Campanian Maastrichtian), the Colombian Western
Cordillera, an oceanic terrane, collided with the South American plate along an active
subduction zone, causing the uplift of the Central Cordillera from south to north. Collision
was oblique (Barrero, 1979). According to Gomez (2001), in the southern part of the
MMV basin the first uplift of the Central Cordillera started in the Maastrichtian and is
evidenced by the Cimarrona formation (outcropping in the southwestern side of the MMV
basin), which consists of a series of conglomeratic fan and braided delta deposits
interfingered with calcareous marine sediments. The source of this sediments were the
igneous rocks of the ancient Central Cordillera (Gomez, 2001). In other areas of the basin,
deformation of the western border of the MMV basin is evidenced by an angular
unconformity that separates Paleocene-age regressive delta or coastal plain facies (Lisama
Formation) from the younger sediments accumulated in a shallow-marine setting (Umir
Formation). Thus a sequence constituted by the Lisama and Umir formations records the
transition from marine to continental conditions.
Coeval with the uplift of the Central Cordillera, a major compressional phase took
place and produced east-verging thrusting. This event also triggered deformation in the
foothills of the present-day Eastern Cordillera (east edge of the MMV basin) with
formation of west-verging thrusts. In addition, inside the MMV basin the highs of Cchira
and La Cira-Infantas were formed, by reactivation of the ancient normal faults (figure 4b
and 4c).
8
Oblique collision of the Western Cordillera also caused development of regional
strike-slip movements in a NE-SW direction. Major faults such as the right-lateral
Palestina, Cimitarra, and Ibague faults cut the western and southern sides of the MMV
basin and modify the previous extensional style. The central and northern parts of the
MMV basin are dominated by transpressional structural deformation triggered by a
restraining step-over of the right-lateral Palestina fault (Feininger, 1970; Barrero and
Vesga, 1976). Feininger (1970) believes that the age of these faults ranges from Late
Cretaceous to pre-Pliocene, which agrees with the age of the collision of the Western
Cordillera terranes against the South American margin.
In Middle Eocene the first compressional phase ceased momentarily, causing
sedimentation above a regional unconformity, known as the Middle Eocene Unconformity,
over the east-vergent thrust belt. The Cachira-Sogamoso and La Cira highs were buried
(figure 4). Nevertheless, these highs kept some relief that allowed the development of two
depocenters next to their flanks where a sequence constituted by La Paz and Esmeraldas
formations accumulated (Suarez, 1996).
During Oligocene time, the Central Cordillera continued to uplift. This caused the
displacement of the depositional axis to the east and the accumulation of the sequence
constituted by the Mugrosa formation (Suarez, 1996).
2.3.3. Miocene - Recent: Broken Foreland (Intermontane) Basin development
By Early Miocene, about 20 Ma, a second major compressional event began as a
consequence of the collision of the Panama-Choco island arc with the northwestern edge of
South America (figure 1). This event is regionally known as the Andean Orogeny and is
9
recorded by the Miocene regional unconformity in the MMV basin. Although late Eocene-
Oligocene deformation had repercussion in the southern part of the Eastern Cordillera, it is
the Andean Orogeny that was responsible for the thick sedimentary pile deposited in the
MMV and for the final configuration of the basin (Dengo and Covey, 1993, Rolon et al.,
2001). A west-vergent thrust belt developed in the western foothills of the Eastern
Cordillera and extended into the MMV basin (figure 4c). The Andean compressional phase
caused high topographic relief that was rapidly eroded and deposited within the MMV
basin. Thus, the basin was covered by Paleogene sediments of the Colorado formation
(Suarez, 1996) sourced from the Central Cordillera and Neogene sediments of the Real and
Mesa formations sourced from both the Central and Eastern cordilleras (Suarez, 1996).
At this stage the MMV basin became an intermontane basin. The most dramatic
deformation was concentrated in the Eastern Cordillera foothills, evidenced by the
presence of imbricated thrust fans, and triangle zones.

3. BACKGROUND ON RIFT BASINS
3.1. Structure
A rift basin is defined as a large elongated tectonic depression bounded by normal
faults that is formed as a zone of rupture in an extensional setting (Burke, 1977; Olsen,
1995). As consequence, normal faulting occurs and the crust is displaced downward along
the fault surfaces creating half-grabens or depressed areas inside the rift. Half-graben
structures are characterized by planar rift-border faults and domino-style intra-rift faults
(McClay, 1998). According to studies on modern rift basins, such as the East African Rift
10
system (Rosendahl et al., 1986; Morley, 1988), the main rifts are markedly asymmetric
with half-graben structures dominant (figure 6). Rosendahl et al. (1986) identified a basic
unit of an arcuate half-graben (figure 7) that is repeated along strike in different patterns of
interaction and overlap.
The simple asymmetric half-graben structure can vary depending on the manner in
which fault overlap is carried out. Change of the polarity of the main faults occurs through
accommodation zones that correspond to structural highs (figures 8 and 9) (Khalil, 1992;
Ebinger, 1989).
The structural geometries of rift basins vary significantly depending on the
mechanical behavior of the pre-existing rocks and of the rocks formed during the
formation of the rift (prerift and synrift packages), the tectonic activity before and after
rifting, and the obliquity of rifting. According with Whithjack and Schlische (2001), some
of the ways pre-existing structural features can affect the geometry of a rift are:
a-) Basins that have salt or thick shale in the prerift or synrift packages are
characterized by forced folds above basement-involved normal faults, diapiric structures,
detached normal faults, and associated fault-bend folds (e.g., southern Suez rift; J eanne
dArc basin, SE Canada).
b-) Basins where one or more major contractional events preceded rifting may
result in normal faults formed from reactivated basement-involved thrust faults (e.g.,
Fundy basin, SE Canada). These faults are commonly low-angle and have large
displacements.
11
c-) Basins produced by oblique rifting are characterized by faults with strike-slip,
normal, and oblique-slip displacement and with multiple trends (i.e., parallel and oblique
to the rift trend) (e.g., Dampier basin, NW Australia).
Kinematics of the normal faults during the opening of the rift affects the thickness
of the synrift deposits. The degree of fanning in the pattern of the synrift deposits is
controlled by the magnitude of rotation of the normal faults (Cartwright, 1992).
The evolution of a rift takes place in three phases: initial aperture, synrift
subsidence, and postrift drift, which are carried out over a relative short period of time, and
very rapid rate of subsidence and sedimentation (Landon, 1994). In their mature stages
new crust of oceanic affinity can be created and rift can evolve either into a passive margin
or an aulacogen (failed rift). Due the scope of this study only the initial aperture and the
synrift subsidence phases will be discussed.
3.1.1. Initial Aperture
Many authors (Sengor and Burke, 1978; Baker and Morgan 1981; and Turcotte and
Emerman, 1983) propose that rift formation processes that are responsible for generation
of extensional strains could be divided in two main classes: active and passive. Active
rifting is related to a hot spot or plume of magma coming from the mantle (Wilson, 1969)
that causes thinning and uplift of the lithosphere, generating tensional stresses. Active
rifting is accompannied by synrift volcanism. The East African rift represents a good
example of active rifting. In passive rifting, fracturing of continental lithosphere as
consequence of tensional stresses, allows hot mantle rocks to penetrate the crust through
weakness zones. According to Miall (1985) most rifts are passive and most of them have
12
originated at divergent plate margins during the breakup of the supercontinent Pangea in
the early Mesozoic (Dietz and Holden, 1970) or during breakup of the supercontinent
Rodinia in the Paleozoic (Dewey and Bird, 1970; Bond and Kominz, 1984; and Moores,
1991).
3.1.2. Synrift Subsidence
Subsidence is probably the most important process that influences the infill of any
basin. In the case of rift basins, subsidence takes place progressively after the initial phase
of aperture whether they are formed by active or passive processes. According with
Dickinson (1974), subsidence of the upper surface of the crust is induced by the following
processes:
Thinning of crust due to stretching, erosion, and magmatic withdrawal.
Thickening of mantle lithosphere during cooling.
Sedimentary and volcanic loading of both crust and lithosphere.
Tectonic loading of both crust and lithosphere.
Dynamics effects of asthenospheric flow.
Crustal densification.
For some authors (e.g. Ensile, 2000; McKenzie, 1978) the main mechanism of
subsidence in rift zones is stretching, although the mechanisms through which this type of
deformation is produced are still not clear enough. Other scientists however, think that
extension is not the only factor that controls subsidence in a rift during the syn-rift stage
but that the sediment load also plays an important role in this process (Miall, 1985; Ziegler,
13
1992), and that the mechanical stretching of the lithosphere and the thickness and thermal
state of the crust are factors than also can influence the structural style of the basin.
3.1.3. Inversion of extensional structures
Many regions have undergone reversal in the sense of vertical movement of the
basement at different stages in their evolution. This reversal is known as inversion when it
happens as the result of a switch in tectonic mode from extensional tectonics to
compressional tectonics such that extensional basins become positive structural features
(positive inversion). Nevertheless, there can be negative inversion when an uplifted region
subsides. The most common type is positive inversion (Holdsworth et al., 1997). Inversion
can arise from several mechanisms. However, compressional/transpressional tectonism
associated with continent collision as a result of plate motions constitutes the most
probable cause for inversion, uplift, and deformation of a extensional hangingwall basin
(Ziegler et al. 1995). Inversion of tensional hanging-wall rift basins located behind a
magmatic arc is the result of acceleration of convergence rates between the colliding plates
and the transmission of compressional stresses into the backarc domain of the overriding
plate (Ziegler et al., 1998). Many aspects of the MMV basin fit this model. Experimental
work of inversion structures shows that thrust and reverse faults commonly use pre-
existing extensional structures (McClay and Buchanan, 1991), especially where the
original structures had a high-angle dip. These normal faults are reactivated as reverse
faults, and as associated contractional fault-bend and fault-propagation folds (Whithjack
and Schlische, 2001). Distinctive characteristics of positive inversion geometries are
thicker synrift strata on the hanging wall of thrust faults, variation of fault-throw with
14
depth (Williams, Powell, and Cooper, 1989; Holdsworth et al., 1997) (figure 10), and
footwall shortcut thrusts (Hayward and Graham, 1989). Shortcuts are formed when a pre-
existing normal fault is sufficiently steep that it inhibits efficient thrusting. Therefore the
new inverted structure has to develop a new fault plane that uses the apex of the old normal
fault to define its pathway (figure 11).
3.2. Stratigraphy
Different factors exert control on the stratigraphy of a rift basin, especially during
its initial apperture or synrift phase. Sedimentation is mainly controlled by subsidence
rates, however, structural asymmetry of the rift and its uplifted flanks also affects the
distribution of sedimentary environments and consequently the distribution of the facies
(Busby and Ingersoll, 1990). As Ravns and Steel (1998) state, structure of the basin exerts
control in accommodation space and sediment supply. During periods of seismic activity,
movements of fault blocks involve hanging wall subsidence and footwall uplift, which
leads to hanging wall deposition and footwall erosion. Footwall uplift in turn influence the
pattern of the rift shoulders acting as barriers to sediment transport (Allen and Allen,
1990).
The nature of the sedimentary fill of a rift basin also depends on deposition of
volcanic flows, which can reach many hundreds of meters in thickness, glacio-eustatic
changes, and climate (Allen and Allen, 1990; Miall, 1985; Ravns and Steel, 1998).
Glacio-eustatic changes control the accommodation space. Climate includes factors such as
temperature, wind level, rainfall and weathering processes that can influence in the
accumulation of deposits such as red beds and evaporites.
15
3.2.1. Early synrift deposits
According with Miall (1985), deposition of the infill sequence in a rift basin
initiates during continental splitting phase of the rift. At this stage, floors of the newly
created grabens remain above sea level, and are caped by nonmarine sediments such as
alluvial, lacustrian, volcanic and/or eolian deposits. In general, rivers flow longitudinally
along the rift whereas, alluvial fans are formed by rivers draining transversely across fault
scarps. In turn, lakes may occupy saddle areas along the rift. At this stage facies show a
markedly asymmetric pattern, which is caused by differential sediment supply and
depositional processes on the two sides of half-grabens, particularly, if most of the material
is supplied from their flanks (Einsele, 2000). Sediments on the steep flank (footwall) of the
rift tend to be coarse grained and form alluvial fans, fan deltas, and debris flows in deeper
water. To Chakrabarti and Mukherjee (1997), presence of these deposits denotes
dislocation periods during opening of the rift. The length of the tectonic slope formed
during extension controls the drainage area of the highlands and drainage area is in turn the
primary control on the size of alluvial fans that are formed across graben margins. Inside
basin lows lake deposits form. Thus, gypsum, halite and other evaporite minerals are often
present.
With continued subsidence, transverse river valleys may become submarine
canyons and alluvial fans may continue to form as submarine fans which are constituted by
clastic turbidites and other gravity flow systems (Miall, 1985). If the basin is filled
predominantly by a through-flowing axial river or by an axial delta lobe in conjunction
with a prograding submarine fan, both the main trunk of the river and the submarine fan
16
tend to shift towards the boundary fault (Einsele, 2000). Shallow marine deposits can also
be derived from the hanging wall. Stacking pattern and geometries of this kind of
sediments depends on the interplay of accommodation in relation with sediment supply
(Ravns and Steel, 1998). Facies of hangingwall on the dip slope includes shorelines and
associated backbarrier environments. If climatic change or significant submergence causes
the submergence of uplands around the half grabens, siliciclastic debris supply is cut off
and then chemical and biochemical sediments accumulate, giving rise to a marine
carbonate regime. Under these conditions, the hangingwall represents a setting where reefs
and their associate facies may develop and finally create a morphology similar to that
along the boundary fault (Einsele, 2000). In addition, any regional or local restriction of
marine circulation under arid climates, causes chemical evaporite deposition, either
subaerial sabkha and subaqueous marine types (Busby and Ingersoll, 1990).
3.2.2. Late synrift deposits
Continued regional subsidence increases water depth creating deep marine conditions.
If the rift reaches mature stage, new oceanic crust is created and a new oceanic stage is
established (Miall, 1985). At the incipient oceanic stage of the basin, accumulation of
clastic sediment may continue under progressively deeper water conditions and
sedimentation reaches the oceanic crust. Three facies assemblages characterize this stage
of basin development: evaporites, marine muds enriched in organic mater, and starved
basin calcareous and pelagic rocks.
17
According to Miall (1985), evaporite deposition can be a consequence of the early
stages of marine inundation when connection with the sea is spasmodic and controlled by
eustatic sea level changes, or from the concentrated brines in deep, isolated basins.
Deposition of black organic shales is due to high organic productivity and restricted
marine circulation. Thus, deep marine rift basins are enriched in organic muds, which
makes them excellent petroleum source rocks.
Where continental margin relief is low, sediment supply is minimal and
sedimentation may be dominated by platform and pelagic carbonates and other fine-
grained, pelagic facies. These sediments rest directly on oceanic crust (Miall, 1985).
The latest synrift stage is characterized by warning or low rates of extensional
faulting, therefore there may be sufficient sediment yield potential to exceed the
accommodation space generating shallow depositional enviroments, and allowing
progradation of coarse siliciclastic systems (Ravns and Steel, 1998).
3.2.3. Synrift unconformities
According with Ravns and Steel (1998) the synrift unconformity represents the
main erosional surface that forms during continental rifting. It develops over fault-blocks
and rift shoulders. In a rift basin this unconformity results from uplift of a fault-block and
lithosphere unloading as consequence of extension and decay of the heat flow. Synrift
unconformities consist of a single erosional surface in the upper areas of a tilted fault-
block, but they can break up into several minor unconformity strands downdip on the
hanging wall of the normal fault. In such cases the individual unconformity strands
represent the effect of successive rift phases and eustatic sea level variations.
18

4. INTERPRETATION/RESULTS
The northern portion of the MMV basin constitutes a monocline striking northeast-
southwest and dipping southeastward. This monocline is cut by normal and reverse faults
with variable vergence that, in general trend northeast-southwest, although there are also
some structures orientated in a northwest-southeast direction. The faults are very steep,
with short strike lengths (10 km on average) and relatively small vertical offset, usually
two-way time of the vertical offset is less than 0.2 sec (approximately 1100 ft using an
interval velocity of 11000 ft/sec as an average value for the study area; average calculated
based on velocity data provided by Ecopetrol) (figures 12 and 13). Many compressional
structures in the MMV basin are cored by high-angle reverse faults (Ecopetrol, 2001). The
geometries displayed by these structures (figures 14 and 15) are comparable with those
described in the experimental models of inverted normal faults developed by McClay and
Buchanan (1991), and in the theorical models included in Williams, Powell and Cooper
(1989) (figure 10). This analogy is used as an element of support to interpret this structures
as inverted normal faults in this work. Other inversion geometries that are also displayed in
the seismic data are thicker rift strata on the hanging wall of thrust faults, variation of fault
throw with depth, positive inversion structures (figure 10), and footwall shortcut thrusts
(figure 11). Good examples of these types of structures are displayed in the seismic line
NC-81-2 (figure 16).
Thicknesses of the J urassic wedge on the hangingwalls of the faults are higher than
in the footwalls and give the seismic reflectors of this unit a divergent character. Even
19
though this pattern can be observed in most of the seismic lines interpreted for this work,
the line BSM-91-2270 (fault at the NW side in figure 16) shows this geometry very
clearly. According to Cartwright (1992), a divergent pattern in seismic reflectors is an
indicator of activity of the faults, with rotational kinematics during the accumulation of the
syn-rift sequence. In contrast to the syn-rift sequence, the post-rift sequence displays even
and parallel reflectors suggesting a non-rotational regime throughout the sedimentation
period (figures 12, 13, 14, 15, and 16).
Most of the inverted faults, and those that still keep the normal displacement, strike
NE-SW although there is a group of NW-striking faults in the northern part of the valley
(figure 17). Dip direction is not the same in all the faults but I consider that the
predominant direction is southeastward (figures 13, 17, 18, and 19). The syn-rift sequence
involved in these structures thickens to the southeast also, the same direction of the
monocline dip. I believe inversion came about from compressional transpressional
deformation along the western and eastern edges of the basin. In the literature the
compressional event is described as having started in Campanian time and continued until
the Recent with major pulses occurring during the Paleogene and Middle Miocene
(Barrero, 1979; Ecopetrol, 2001; Gomez, 2001; Rolon et al., 2001). However, the inversion
in the study area seems to have happened between Late Cretaceous (Campanian-
Maastrichtian) and Paleocene, since most of the inversion structures are buried by the
Tertiary unconformity (figures 12, 13, and 16). Nevertheless, there are some structures that
do cut the unconformity, indicating that in these cases inversion continued later. In the line
NC-81-2 (figure 16), the inverted structure located at the southeast side of the line is buried
20
by Eocene rocks (age data taken from Santa Catalina 1 well located to the north of the
line) (figure 2). The unconformity here shows minor relief. In contrast, the inverted fault
on the northwest of the line cuts the Tertiary unconformity. Onlap of the strata on the
flanks of the inverted high shows that deformation was ongoing during Eocene time.
Transpression caused by the oblique collision of the Western Cordillera, mainly
developed during Paleocene, may have added a strike-slip component to the reactivation
process. North-east trending faults that outcrop in the Central Cordillera west of the study
area have been interpreted as right-lateral strike-slip faults (figure 3).
4.1. Syn-rift stratigraphy
Infill of the graben represents a 2nd order tectono-sequence that is conformed by
the J ordn, Girn, Santos, Cumbre, Rosablanca, Paja and Tablazo formations. This
sequence is a second order cycle that corresponds with the Giron-Rosablanca, Paja
Arenoso and Paja Tabalazo defined in the sequence-stratigraphy model proposed by Rolon
and Carrero (1995) and the sequences K-1, K-2, and K-3 described in Ecopetrol (2001).
The upper sequence boundary depicted in these documents is supported on paleontological
data published in Etayo (1994) and Garzon (2000), the lower boundary does not have
paleontological support but can be followed in the seismic data as I will describe later. The
lowstand systems tract of this sequence is conformed by the J ordn, Girn, and Santos
formations, the transgressive systems tract by the Cumbre and Rosablanca formations and
the highstand systems tract by the Paja and the Tablazo formations. Rolon and Carrero
(1995) present an extended facial analysis of the units mentioned above for the
21
surrounding area to the Los Cobardes anticline, located to the south of the study area
(figure 3).
The early synrift deposits, represent the lowstand systems track of the sequence.
This rocks were accumulated during the splitting phase of the rift. At the base of this
interval there is a regional unconformity that can be recognized from truncations of the
reflectors and change in the character of the seismic data, from irregular reflectors to more
defined and continuous reflectors (figures 5b, 12, 13, 14, and 15). The early synrift
sequence is characterized by fluvial and lacustrine facies of the Girn, J ordn and Santos
formations. According to field observations described in Rolon and Carrero (1995) and
Ecopetrol (2001), the facies of the J ordn, Girn, and Santos formations constitute
sandstone and conglomerate facies, alternating with red, gray and green mudstone.
Channels and lenses are the predominant sedimentary structure, with associated internal
cross bedding. At the top of the sequence the sedimentary structures turn to flasser and
wavy bedding. In the subsurface this sequence presents a seismic display characterized by
irregular, discontinuous reflections, rotated by normal-fault blocks (figure 12). In addition,
the electrofacies displayed by the SP and GR logs present a funnel shape which is and
indicator of fluvial deposits (point bars) (figure 20).
The early synrift deposits form an asymmetric sedimentary wedge that thickens
toward the southeast (figure 21). This asymmetric pattern can be related to the geometry of
half-grabens previously described in the background section. The upper boundary of this
interval is a sharp contact (at the top of the Arcabuco formation) that represents a
ravinement surface (Rolon and Carrero, 1995).
22
The upper level that is cut by normal faults represents the late synrift deposits and,
in turn, the transgressive and highstand systems tracts of the synrift sequence (Rolon and
Carrero, 1995). This interval is conformed by limestones and siliciclastic shales of the
Basal Lime Group.
Facies progressively change from evaporites and packstones, accumulated under
tidal conditions, to wackstones, mudstones, and shales accumulated on a platform that was
progressively deepening through a regional subsidence period (Rolon and Carrero, 1995,
Ecopetrol, 2001). The thickness of this interval is about 1000 feet (0.1 seconds) and is kept
more or less homogeneous throughout the study area (figure 22), however it increases
substantially to the south where there are four well-differentiated units, the Cumbre,
Rosablanca, Paja, and Tablazo formations. The seismic character of the Basal Lime Group
in the study area consists of continuous reflectors with gradual changes in thickness. The
facies associated with this interval are: evaporites in the Cumbre Formation; calcareous
sandstones, mudtstones, and wakestones in the Rosabalanca formation; shales in the Paja
Formation and mudstones in the Tablazo Formation (figure 23). In the eastern part of the
Middle Magdalena Basin and in the Eastern Cordillera, the Rosablanca and Paja
formations have been reported to include evaporite facies (Etayo, 1968; Rolon and Carrero,
1995). In addition, descriptions of the lower Albian presented in Ecopetrol (2001), mention
presence of rocks characterized by coarse-grained siliciclastic rocks originated in
continental environment, bordering the Santander Massif and adjoining to the
Bucaramanga-La Salina fault system (figure 24).
23
According to paleontological data presented in Rolon and Carrero (1995) and
Garzon (2000), the Basal Lime Group presents an heterochronous character along the
basin. South of the study area the base of this unit corresponds to the Rosablanca formation
dated as Valanginian while in the study area, in the northern part of the basin, the
Rosablanca formation has been omitted and the age reported for the base of Basal Lime
Group is early Albian (Garzon, 2002). Since the age of the rocks that directly underlay the
Basal Lime Group is Late J urassic (Geyer, 1973; Barrero, pers. commun.), the absent
section can be interpreted as a hiatus caused by the continuous exposure and erosion of the
area.
At the top of the Basal Lime Group, a regional unconformity is developed: the
synrift unconformity. This unconformity can be recognized in the field in the area of the
Los Cobardes anticline by an irregular contact with the upper strata, which represents an
erosion and subaerial exposure surface (Rolon and Carrero, 1995). According with Roeder
and Chamberlain (1995) this unconformity is dated as Early Aptian and represents the first
thermal decay phase that marks the end of the extension. Thermal subsidence can cause a
compensating uplift of rift margins (Van der Beek, 1995 cited in Sarmineto, 2001). The
thermal subsidence lead to the definitive establishment of a marginal-sag type basin.

5. DISCUSSION
The monocline structure that constitutes the study area dips southeastward, in the
same direction towards which the synrift sequence is thickens. To the east of the area, are
the La Salina, and Bucaramanga fault-systems that represent the frontal fault of the Eastern
24
Cordillera foothills. Their general strike is northwest-southeast. Even though these features
currently constitute a thrust and a strike slip fault-systems, authors in previous works (e.g.
Sarmiento, 2001) have interpreted them as a former normal-fault system of western
vergence (see figure 24), that was inverted during the Tertiary. Faults to the west of the La
Salina and Bucaramanga fault-systems are typically inverted normal faults.
To the west of the Bucaramanga fault-systems the J urassic synrift wedge is up to
2000 m thick while in the Santander Massif, located to the east of this structure, the
thickness of the J urassic sequence is dramatically reduced up to 200 m on average,
according with Etayo and Laverde, ed. (1985). This fact in addition with the facies
distribution of the Lower Cretaceous in the study area (section 4) are reasons to interpret
the Santander Massif as a J urassic-Early Cretaceous paleohigh (Laverde, 1985; Rolon and
Carrero, 1995, Ecopetrol, 2001).
The predominant strike direction of the normal faults interpreted in this work is
northeast-southwest and their current sense of slip is variable as the result of Tertiary
inversion. However, the geometry of the J urassic wedge, thickening to the southeast, and
the presence of a high to the east of the La Salina and Bucaramanga fault-systems makes
the geometry of the study area to be comparable with the model of rift proposed by
Rosendahl et al. (1986). In this model the synrift sequence thickens towards the main
normal fault (see sketch A3-A3 in figure 7). Therefore, most of the secondary normal
faults of the northern MMVB that dip towards the La Salina and Bucaramanga fault-
systems are antithetic to the master fault. According with this model it is also possible to
25
interpret that the former vergence of the La Salina and Bucaramanga fault-systems was to
the southwest (figures 25 and 26).
South of the study area is the Cambao-Bituima fault-system. Although this
structure was not analyzed in this study, interpretations presented in Sarmiento (2001)
(figure 24) and Cediel, Barrero, and Caceres (1998) are taken in to account for the final
model proposed in this work.
The Cambao-Bituima fault-system trends northeast-southwest. It is currently a
northwest vergent thrust system, however, during the rift opening, this system represented
a normal fault-system verging toward the southeast. Eastward of the Cambao-Bituima fault
system, the thickness of the J urassic wedge increases considerably in the areas of the
Portones and Los Cobardes anticlines (Rolon and Carrero, 1995), while to the west of the
fault system, the same unit is pinches out at the Central Cordillera (Cediel, Barrero and
Caceres, 1998) (figure 26).
I conclude that the study area and the southern part of the MMV basin present the
geometries of two half-graben structures with opposite polarities. This pattern can be
compared with the pattern observed by Rosendahl et al. (1986), on the active East African
Rift system (figures 25 and 8). In the case of the MMV basin, the master normal faults that
limited the half grabens corresponded with the paleo-La Salina fault-system in the study
area and the paleo-Cambao-Bituima fault-system in the southern MMV basin (figure 24
and 26). The half-grabens overlap and switc their polarities through an accommodation
zone in the same way that is shown in the models proposed by Rosendahl et al. (1986),
26
Khalil (1992), and Ebinger (1989) (figure 9). This accommodation zone corresponds with
the area where the La Cira high is currently located (figure 26).

6. CONCLUSIONS
The northern part of the MMV basin constitutes a monocline structure dipping to
the southeast. Its general configuration is characterized by a synrift sequence
thickening to the southeast also, and a set of normal faults with variable vergence
that are often antithetic to the paleo-La Salina and paleo-Bucaramanga fault-
systems.

The synrift sequence is bounded at the base and at the top by regional
unconformities that can be followed easily on the seismic data.

During J urassic and early Cretaceous the MMV basin went through a rift stage.
Secondary normal faults striking northeast-southwest were developed in the interior
of the MMV basin during this period. The former vergence of these faults at the
time of the rifting was most commonly down-to-the-southeast.

Most of the faults remaining from the rift stage were inverted after the
compressional tectonics that started at the Late Cretaceous (Campanian).
Additionally, a strike-slip component was added mainly during the Paleocene time.
These further deformations came about from compression- transpression along the
27
western and eastern edges of the basin through the oblique collision of the Western
Cordillera.

The geometry of the northern part of the MMV basin during J urassic time can be
interpreted as a half graben that follows the same pattern that Rosendahl et al.
(1986) described based on observations on the active East African Rift system.

From a more regional point of view, during J urassic time the MMV basin consisted
of a pair of half-grabens that switched their polarity through an accommodation
zone. The main normal faults that bounded the half graben structures correspond
with the paleo-La Salina-Bucaramanga and paleo-Cambao-Bituima fault systems
and the accommodation zone where the main faults were overlapping corresponded
with the paleo-La Cira high.

Understanding of the J urassic structures helps to understand better the geometry
and genesis of the Tertiary traps in the MMV basin as well as the potential areas of
hydrocarbons kitchens and migration pathways.
28
REFERENCES CITED
Allen, P., and Allen, J., eds., 1990, Basin Analysis, Principles and Applications.
Blackwell Scientific Publications, Oxford, 451 p.
Baker, B.H., and Morgan, P., 1981, Continental rifting: progress and outlook. Eos, Trans.
Am. Geophys. Union, 62, p 585-586.
Barrero, D., and Vesga, C.J., 1976, Mapa Geolgico del Cuadrngulo K-9 Armero y
parte sur del J -9, la Dorada, 1:100.000. INGEOMINAS, Bogot, Colombia.
Barrero, D., 1979, Geology of the Central Western Cordillera, west of Buga and
Roldanillo, Colombia. Publicaciones Geolgicas Especiales INGEOMINAS no.4,
p.1-75.
Barrero D., J. Alvarez, and T. Kassem, 1969, Actividad gnea y tectnica en la
Cordillera Central durante el Meso-Cenozoico: Boletn Geolgico INGEOMINAS,
no.1-3; p. 145-173.
Bond, G.C., and Kominz, M.A., 1984, Construction of tectonic subsidence curves for
early Paleozoic miogeocline, southern Canadian Rocky Mountains: implications for
subsidence mechanisms, age of break up, and crustal thining. Bull. of the
Geological. Society of America, 95, p. 155-173.
Burke, E., 1977, Aulacogens and continental break-up. Ann. Rev. Earth Planet Sci., 5, p.
371-396.
Busby, C., and Ingersoll, R. eds., 1995, Tectonics of Sedimentary Basins. Blackwell
Science, Cambridge, Mass., USA, 579 p.
29
Cartwright, J., 1992, The kinematic evolution of the Coffee Soil Fault: Roberts, A.,
Yielding, G., & Freeman, B. (eds). The Geometry of Normal Faults. Special
Publication of the Geological Society of London, p. 29-40.
Cediel, F., D. Barrero, and C. Cceres, 1998, Seismic expresion of structural styles in
the basins of Colombia: Six volumes, prepared for Geotec-Ecopetrol, edited by
Robertson Research, London.
Cooper, M.A., Addison, F.T., Alvarez, R., Coral, M., Graham, R.H., Hayward, A.B.,
Howe, S., Martinez, J., Naar, J., Peas, R., Pulham, A.J., and Taborda, A.,
1995, Basin Development and Tectonic History of the Eastern Cordillera, Llanos
Basin and Middle Magdalena Valley, Colombia. AAPG Bull., V.19, p. 1421-1443.
Chakrabarti, M K, and Mukherjee, B. K., 1997, Sedimentation and hydrocarbon
prospects of intracratonic Gondwana rift basins, peninsular India. Indian J ournal of
Petroleum Geology, vol.6, no.2, p.60-82.
Coletta, B., Herbard, F., Letouzey, J., Werner, and Rudkiweicz, J., 1990. Tectonic
Style and crustal structure of the Eastern Cordillera (Colombia) from balanced
cross section. ECOPETROL internal report, Bogot.
Dengo, C.A., and Covey M. C., 1993, Structure of the Eastern Cordillera of Colombia;
Implications for trap styles and regional tectonics. AAPG Bull., V.77, p. 1315-
1337.
Dewey, J.C., and Bird, J.M., 1970, Mountain belts and the new global tectonics. J ournal
of Geophysical Research, Vol. 75, p. 2625-2647.
30
Dickinson, W.R., 1974, Plate tectonics and sedimentation. In Tectonics and Sedimentation
(Ed. By W.R. Dickinson), 1-27, spec. Publ. SEPM, 22, Tulsa, Oklahoma.
Dietz, R.S., and Holden, J.C., 1970, Reconstruction of Pangaea: breakup and dispersion
of continents, Permian to present. J ournal of Geophysical Research, Vol. 75, p.
4939-4956.
Ebinger, C.J., 1989, Geometric and kinematic development of border faults and
accommodation zones, Kivu-Rusizi rift, Africa. Tectonics, Vol. 8, No. 1, p. 117-
137.
Ecopetrol, 2001, Proyecto Evaluacin Regional Cuenca Valle Medio Del Magdalena -
Cordillera Oriental, Colombia. Internal report.
Einsele, E., 2000, Sedimentary Basins: evolution, facies, and sediment
Budget. Berlin, Germany: Springer Verlag.
Etayo, F., 1968. El Sistema Cretcico en la regin de Villa de Leiva y zonas prximas.
Geologa Colombiana No. 5 p. 5-74. Universidad Nacional de Colombia. Bogot.
Etayo, F., Renzoni, G., and Barrero, D., 1969, Contornos sucesivos del mar Cretcico en
Colombia. Memoria Primer Congreso Colombiano de Geologa, Universidad
Nacional de Colombia, Bogot, p. 217-252.
Etayo, F., Barrero, D., Lozano, H., Espinosa, A., Gonzales, H., Orego, A., Zambrano,
F., Duque, H., Vargas, R., Nuez, A., Alvarez, J., Ropan, C., Ballesteros, I.,
Cardozo, E., Forero, H., Galvis, N., Ramirez, C., and Sarmiento, L., 1983.
Mapa de Terrenos Geolgicos de Colombia. Publicaciones Especiales
INGEOMINAS no. 14, p.1-235.
31
Etayo F. and Laverde F., ed., 1985, Proyecto Cretcico Contribuciones. Publicaciones
Geolgicas Especiales INGEOMINAS no.16, 200 p.
Fabre, A., 1983, 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. 21-21.
Feininger, T., 1970, The Palestina fault. Geol. Soc. America Bull., V.81, p.1201-1216.
Garzon, M.A., 2000, Reporte Palinolgico pozos Esperanza-2, Catalina-1, Totumal-
6,Cascajales-1, Guineal-1, Infantas-1613 y Morales-1, VMM. Reporte de
Bioestratrigrfica para Ecopetrol. Bogot.
Geyer, O., 1973, Das Prakretazische Mesozoikum von Kolumbien (The Precretaceous
Mesozoic of Colombia). Geologisches J ahrbuch V5, 1-156. 5 lam. Hannover.
Gomez, E., 2001, Tectonic controls of the Late Cretaceous to Cenozoic sedimentary fill of
the Middle Magadalena Valley Basin, Eastern Cordillera and Llanos Basin,
Colombia (Ph.D. Thesis). Department of Earth and Atmospheric Sciences, Cornell
University, v. 1, 2, 619 p.
Hayward, A., and Graham, R., 1989, Some geometrical characteristics of inversion:
Cooper, M., & Williams, G. (eds) Inversion Tectonics. Special Publication of the
Geological Society of London, p. 17-39.
Holdsworth, R., Butler, C., and Roberts, A., 1997, The recognition of reactivation
during continental: J ournal of the Geological Society of London, V.154, p. 73-78.
Khalil, M., 1992, Field Guide to the Gulf of Suez rift. AAPG field trip.
32
La Luna Oil Co., 2001, Geologic map with oil fields and wells of the Middle Magdalena
Valley and Eastern Cordillera.
Landon, S., 1994, Summary, in Landon, S., ed. Interior Rift Basins: AAPG Memoir
59, Tulsa, U.S.A., p. 259-270.
Laverde, F., 1985, La Formacin Los Santos: un deposito continental anterior al ingreso
marino del Cretcico. In Etayo F. and Laverde F., ed. Proyecto Cretcico
contribuciones. Publicaciones Geolgicas Especiales INGEOMINAS no.16, p.24.
McClay, K., and Buchanan, P., 1992, Thrust faults in inverted extensional basins:
McClay, K. (ed): Thrust tectonics. Chapman & Hall, London, United Kingdom
(GBR), p. 93-104.
McClay, K., 1998, Tectonic Regimes and fault systems: Structural Geology for Petroleum
exploration. Short course. Egham, Surrey, United Kingdom, 602 p.
McKenzie, D., 1978, Some remarks on the development of sedimentary basins. Earth and
Planetary Science Letters 40, p. 25-32.
Miall, A., 1985. Principles of Sedimentary Basin Analisys. Springer-Verlag,
New York, 490 p.
Mojica, J., Macia, C., and Colmenares, F., 1984. Consideraciones sobre la importancia
de la paleogeografa y las reas de aporte pre-Cretcicas para la prospeccin de
hidrocarburos en el Valle Superior del Magdalena, Colombia. -Mem. 1er Cong.
Col. Petrol, Oct 23-26
33
Mojica, J., and Dorado, J., 1987, el J ursico anterior a los movimientos intermalmicos en
los Andes Colombianos. Bioestratigrafa de los Sistemas Regionales del J ursico y
Cretcico de America del Sur, Mendoza, p. 49-110.
Moores, E.M., 1991, Southwest U.S.-East Antarctic (SWEAT) connection; a hypothesis.
Geology, Vol.19, no.5, p.425-428.
Morales, L. G. et. al., 1958. General geology and oil occurrences of Middle Magdalena
Valley, Colombia. In L. G. Weeks ed. A symposium conducted by A.A.P.G.:
Habitat of Oil, p. 641-695.
Morley, K., 1988, Variable Extension in Lake Tanganyika. Tectonics, Vol.7, p. 785-801.
Olsen, K. H., 1995. Continental rifts: evolution, estructure, tectonics. Elsevier,
Amsterdam, 490 p.
Pindell, J., and Tabbutt, K., 1995, Mesozoic Cenozoic Andean paleogeography and
regional controls on hydrocarbon systems, in A. J . Tankard, R. Suarez, and H. J .
Welsink, Petroleum basins of South America: AAPG Memoir 62, p. 101-128.
Ravns, R., and Steel, R.J., 1998, Architecture of marine rift-basin successions. AAPG
Bull., V. 82, p. 110-146.
Roeder, D., and Chamberlain, R., 1995. Eastern Cordillera of Colombia: J urassic-
Neogene crustal evolution, in A. J . Tankard, R. Suarez, and H. J . Welsink,
Petroleum basins of South America: AAPG Memoir 62, p. 633-645.
Roln L.F., and M. Carrero, 1995, Anlisis Estratigrfico de la Seccin Cretcica
aflorante al oriente del Anticlinal de Los Cobardes entre los municipios de
34
Guadalupe, Chima y Contratacin, Departamento de Santander: Trabajo de grado,
Depto. de Geociencias, Universidad Nacional de Colombia, Bogot, 125 p.
Roln, L.F., Lorenzo, J.M., Lowrie, A., and Barrero, D., 2001, Thrust, kinematics and
Hydrocarbon migration in the Middle Magdalena Basin, Colombia (S.A.). 21
st

Annual GCS-SEPM Conference. Houston, December 2001.
Rosendahl, R.R., Reynolds, DJ., Lorber, P.M., Burgess, C.F., McGill J., Scott, D.,
Lambiase, J.J., and Derksen S.J., 1986, Structural expressions of rifting: lessons
from Lake Tanganyika, Africa. In: Frostick L.E. et al. (Eds.) Sedimentation in the
African Rifts. Geol. Soc. Lond. Spec. Publ. 25, p. 29-43.
Sarmiento, L.F., 2001, Mesozoic rifting and Cenozoic basin inversion history of the
Eastern Cordillera, Colombian Andes. Inferences from tectonic models. Ph.D.
Thesis, Vrije Universitteit, Amsterdam, 295 p.
Schamel, S., 1989, Middle and Upper Magdalna basins, Colombia, in K.T. Biddle, ed.,
Active margin basins: AAPG Momoir 52, p. 283-302.
Sengor, A. and Burke, K., 1978, Relative timing of rifting and volcanism on Earth and its
tectonic implications. Geophysics Research Letters, 5, p. 419-421.
Suarez, M., 1996. Facies analysis of the upper Eocene La Paz Formations, and regional
evaluation of the post middle Eocene stratigraphy, norther Middle Magdalena
Valley Basin. Unpublished MSc, Thesis, University of Colorado, Boulder, 88 p.
Toro, J., 1990. The termination of the Bucaramanga Fault in the Cordillera Oriental,
Colombia. M. S. thesis. Department of Geosciences University of Arizona.
35
Toro, J., Roure, F., Bordas-Le Floch, N., Le Cornec-Lance, S., and Sassi, W., In press,
Thermal and Kinematic Evolution of the Eastern Cordillera fold-and-thrust-belt,
Colombia, in Deformation, fluid flow and reservoir, in Roure, F. and
Swennen, R., eds. Appraisal in foreland fold-and-thrust belts: AAPG Memoir,
Tulsa, U.S.A., 40 p., 20 figs., 8 plates.
Turcotte, D., and Emerman, S.H., 1983. Mechanisms of active and passive rifting.
Tectonophysics, Vol.94, p. 39-50.
Van der Beek, P., 1995. Tectonic evolution of continental rifts, inferences from numerical
modeling and fission track thermochronology. Ph.D. Thesis, Vrije Universitteit,
Amsterdam, 232 p.
Villamil, T., 1999, Campanian-Miocene tectonostratigraphy, depocenter evolution and
basin development of Colombia and western Venezuela. Palaeogeography,
Palaeoclimatology, Palaecology, 153, p. 239-275.
Whithjack, M., and Schlische, R. W., 2001, Structural styles of rift basins. Annual
Meeting Expanded Abstracts - American AAPG Bull., Vol.85, No. 6, p.217.
Williams, G.D., Powell, C.M., and Cooper, M.A., 1989, Geometry and kinematics of
inversion tectonics. Geological Society Special Publications, Vol.44, p.3-15.
Wilson, J.T, 1969, Aspects of the different mechanics of ocean floors and continents.
Tectonophysics, Vol.8, No.4-6, p.281-284.
Ziegler, P.A., 1992, Geodynamics of rifting and implications for hydrocarbon habitat. In
Ziegler, P.A. ed. Geodynamics of rifting; Volume III, Thematic discussions.
Tectonophysics, Vol.215, No.1-2, p.221-253.
36
Ziegler, P.A, Cloetingh, S.,and VanWees, J., 1995, Dynamics of intra-plate
compressional deformation; the Alpine Foreland and other examples.
Tectonophysics, Vol.252, No.1-4, p.7-59.
Ziegler, P.A., VanWees, J., and Cloetingh, S.A., 1998 Mechanical controls on collision-
related compressional intraplate deformation. In Ziegler, P.A. ed. Geodynamics of
rifting; Volume III, Thematic discussions. Tectonophysics, Tectonophysics,
Vol.215, No.1-2, p.103-129.

37
Nazca
Plate
6 cm/y
Caribbean Plate
1-2 cm/y
80W 75W 70W
10N
5N
C
e
n
t
r
a
l

C
o
r
d
i
l
l
e
r
a
W
e
s
t
e
r
n

C
o
r
d
i
l
l
e
r
a
R
S
L
l
a
n
o
s

B
a
s
i
n
Maracaibo
Basi n
V
e
n
.
C
o
l
.
P
F
L
S
F
S
M
M
M
V
B
B
F
E
a
s
t
e
r
n

C
o
r
d
i
l
l
e
r
a
C
a
r
i
b
b
e
a
n

S
e
a
N
1
,
0
0
0
,
0
0
0
1
,
0
5
0
,
0
0
0
1,450,00
1,400,000
1,350,000
1,300,000
Bog
Bog - Bogota
RS - Romeral Suture
PF - Palesti na Faul t
BF - Bucaramanga Fault
SM - Santander Massi f
LSF - La Sali na Fault
MMVB - Middle Magdalena
Valley Basin

Middle
Magdalena
Valley
Basin
Study Area
Figure 1. Location of the study area on a tectonic map of the Northern Andes (modified from Toro et al., in press)
38
1
,
0
0
0
,
0
0
0
1
,
0
5
0
,
0
0
0
1,450,00
1,400,000
1,350,000
1,300,000
N
Seismic Line
B-89-1420 B-89-1420
Wel l

0 50Km
Norean -1
Crisol-2
Trigos-1
Totumal-6
Gironda-1
Bosques- 3
Figure 23
Wells discussed
in the text
Well
Figure 15
Figure 13
Figure 12
Figure 16
Figure 14
Santa Catalina-1
Catalina-1
Seismic Line

Figure 2. Distribution of wells and 2D seismic data used for this thesis.
39

Figure 3. Geology map of the Middle Magdalena Valley showing the main faults and the
distribution of the units involved in this study (Sources: Ecopetrol,2001; La Luna Oil
Co., 2001).
40


La Salina Fault

Figure 4. Cross sections through the northern part of the Middle Magdalena Valley Basin (see location on figure 3)
(adopted from Ecopetrol, 2001).

41











































Figure 5a. Generalized stratigraphic column of the study area showing the main boundaries
followed in the seismic lines interpreted for this thesis.
42
REAL
COLORADO
MUGROSA
ESMERALDAS
LA PAZ
LISAMA
C
O
N
T
I
N
E
N
T
A
L

UMIR
LA LUNA
SIMITI
PAJ A
TABLAZO
ROSABLANCA
CUMBRE
ARCABUCO -
LOS SANTOS
GIRON
CONTINENTAL
Fluvial systems
to coastal deposits
under tidal influence
NEOGENE
PALEOGENE
Maastrichtian
Campanian
Santonian
Coniacian
Turonian
Cenomanian
Albian
Aptian
Barremian
Hauterivian
Valanginian
Berriasian
UPPER
J
U
R
A
S
S
I
C

C
R
E
T
A
C
E
O
U
S

T
E
R
T
I
A
R
Y

QUATERNARY MESA
PERIOD
EPOCH / AGE
FORMATION LITOLOGY
DEPOSITIONAL
ENVIRONMENT
MARINE
Delta front
deposits
MARINE
Outer to
Middle shelf
MARINE
Inner to
middle shelf
MARINE
Inner to
middle shelf.
Subtidal zone
MARINE
Middle shelf,
storm dominated
towards the top
MARINE
Calcareous platform
Sabka tidal flats
J
U
R
A
S
S
I
C

W
E
D
G
E

TRANSITIONAL
(Swamps)
B
A
S
A
L

L
I
M
E


O
U
P

G
R
M
I
D
D
L
E

A
N
D

U
P
P
E
R

E
T
A
C
E
O
U
S

C
R
T
E
R
T
I
A
R
Y


Seismic Character
LEGEND

Tertiary
Mid. & Up. Cretaceous
Basal Lime Group
Jurassic Wedge
100 ft

Figure 5b. Seismic character and the GR, SP, and resistivity responses of the units studied in this thesis (see figure 5a). The curves of
the well Bosques-3, located to the south of the study area, are taken as a typical log for the middle and central sectors of the
MMV basin.
43








Figure 6. Isopach maps in two travel time for acoustic units in the East African Rift
System, showing half-graben depocenters and the main rift borders faults
(adopted from Rosendahl et al., 1986).
44







Case of the MMV basin




Figure 7. Asymmetric half-graben profiles generated by different systems of antithetic
and synthetic faults (adopted from Rosendahl et al., 1986).

45











Figure 8. Different geometries of asymmetric half-graben structures as a result of the
change in polarity of the main faults (adopted from Rosendahl et al., 1986).

46



Figure 9. Accommodation zones, Gulf of Suez (adopted from Khalil, 1992).

47


C
B
A
A
B
C
B
A
A
B
C

1.
2.
Figure 10. Schematic diagram of a positive inversion structure according with Williams,
Powell and Cooper (1989).






uplift
shortcut

Figure 11. Model for development of a shortcut thrusting (modified after Hayward and
Graham, 1989).

48




























W E
Flower
Structure
LEGEND
Tertiary
Mid. & Up. Cretaceous
Basal Lime Group
Jurassic Wedge
A
B
Km 1
3.50
3.00
2.50
2.00
1.50
1.00
0.50
0 0.0
ms
4.00


Figure 12. Interpretation of the seismic line BSM91-2270 (see figure 2 for location).
49








N S
1 Km
ms
0.0
















LEGEND
Tertiary
Mid. & Up. Cretaceous
Basal Lime Group
Jurassic Wedge
0
0.5
0
1.0
0
1.5
0
2.0
0


Figure 13. Interpretation of the seismic line B89-1370 (see figure 2 for location).
50






LEGEND
Tertiary
Mid. & Up. Cretaceous
Basal Lime Group
Jurassic Wedge
1 Km
E
Inversion structure
W
ms
0.00
0.50
1.00
1.50
2.00
2.50
3.00
3.50
4.00


Figure 14. Interpretation of the seismic line BSM91-2020E (see figure 2 for location).
51






W E
1 Km
LEGEND
Tertiary
Mid. & Up. Cretaceous
Basal Lime Group
Jurassic Wedge
Inversion structure
ms
0.00
0.50
1.00
1.50
2.00
2.50
3.00
3.50
4.00


Figure 15. Interpretation of the seismic line BSM91-1420 (see figure 2 for location).



52






























1.50
2.00
2.5
0
3.0
0
3.50
4.00
4.50
1 Km
ms
LEGEND
Tertiary
Mid. & Up. Cretaceous
Basal Lime Group
Jurassic Wedge
Shortcut
Positive
Inversion
structure
SE NW
Figure 16. Interpretation of the seismic line NC-81-2 (see figure 2 for location).
53




































































20 Km
Norean -1
Crisol -2
Trigos -1
Santa Catalina-1
Gironda-1
Bosques-3
Totumal -6
Figure 15
Figure 16
F
i
g
u
r
e

1
3

Figure 14
Figure 12
1,300,000
1,350,000
1,400,000
1,450,00
1,050,000 1,000,000





milliseconds


Figure 17. Structural map in time at the base of the J urassic wedge, wells in red are discussed in the text.

54







































































20 Km
Norean -1
Crisol -2
Trigos -1
Totumal -6
Santa Catalina-1
Gironda-1
Bosques-3
Figure 15
Figure 16
F
i
g
u
r
e

1
3

Figure 14
Figure 12
1,300,000
1,350,000
1,400,000
1,450,00
1,050,000 1,000,000



milliseconds

Figure 18. Structural map in time at the top of the J urassic wedge.

55







































































20 Km
Figure 15
Figure 16
Norean -1
Crisol -2
Trigos -1
Totumal -6
Santa Catalina-1
Bosques-3
F
i
g
u
r
e

1
3

Figure 14
Figure 12
Gironda-1
1,300,000
1,350,000
1,400,000
1,450,000
1,050,000 1,000,000



milliseconds

Figure 19. Structural map in time at the base of the Basal Lime Group.

56


Bosques-3 Norean-1


ft
f
t

Catalina-1

















ft






Figure 20. Electrofacies of the Arcabuco-Los Santos formations in the study area (see figure 2 for location of wells).
57


20 Km
Bosques-3
Norean -1
Gironda-1
Santa Catalina-1
Totumal -6
Trigos -1
Crisol -2
1,300,000
1,350,000
1,400,000
1,450,000
1,050,000 1,000,000









































milliseconds

Figure 21. Isopach map (in time) of the J urassic wedge.
58
20 Km
Bosques-3
Norean -1
Gironda-1
Santa Catalina-1
Totumal -6
Trigos -1
Crisol -2
1,300,000
1,350,000
1,400,000
1,450,000
1,050,000 1,000,000










































milliseconds

Figure 22. Isopach map (in time) of the Basal Lime Group.
59

Cumbre
Rosablanca
Paja
Tablazo

1000
Lower Tertiary
Vertical Scale

Middle and Upper Cretaceous

Basal Lime Group
J urassic

Figure 23. Stratigraphic cross-section along the study area based on GR and resistivity logs of wells that reach the top of the J urassic.
Distance between wells in Km. Llanito-1 and Infantas 1613 are located south of the study area.
60































Serrania de
Perija
Bogot
Villavicencio
Medellin
Bucaramanga
Ibague
Nei va
0
100 200 Km
Alluvial fan and
fluvial deposits
Littoral to inner
shelf deposits
Shallow marine
deposits
Deep marine
deposits
Magdalena Valley
Boundary
VENEZUELA
COLOMBIA
LEGEND
1400 1300 1200
1000 1100 900 800
1400
1300
1200
1100
1000
900
800





Figure 24. Paleo-geographic model of the Magdalena Valley and the Eastern Cordillera
for Valanginian-Aptian time (adopted and modified from Sarmiento, 2001).



61




























Half graben model proposed by Rosendahl et al (1986).

Middle Magdalena Basin



Figure 25. Comparison of the geometry of the rift of th MMVB with Model proposed by Rosendahl et al. (1986) for the East African
rift system.

62



Figure 26. Model of the J urassic structural geometry of the Middle Magdalena Valley
Basin proposed in this thesis. The master faults (in purple) represents the
boundaries of the half grabens inside the rift structure that extended farther
than the MMV basin. The accommodation zone, where the faults are
changing their vergence forms a high that matches with La Cira high, where
one of the most important oil fields of Colombia is located.
63

Vous aimerez peut-être aussi