0% found this document useful (0 votes)
3 views40 pages

M108_Chapter16

The document discusses the integrated structural and basinal analysis of the Cesar-Rancheria Basin in Colombia, focusing on its tectonic history and petroleum systems. It outlines the methods used for the study, including seismic data interpretation and crustal structure analysis, and identifies significant tectonic events that have shaped the basin. The findings highlight the influence of tectonic interactions between the Caribbean and South American plates on the basin's sedimentation and petroleum potential.

Uploaded by

fourtime20
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
3 views40 pages

M108_Chapter16

The document discusses the integrated structural and basinal analysis of the Cesar-Rancheria Basin in Colombia, focusing on its tectonic history and petroleum systems. It outlines the methods used for the study, including seismic data interpretation and crustal structure analysis, and identifies significant tectonic events that have shaped the basin. The findings highlight the influence of tectonic interactions between the Caribbean and South American plates on the basin's sedimentation and petroleum potential.

Uploaded by

fourtime20
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 40

16

Sanchez, Javier and Paul Mann, 2015, Integrated structural and basinal analysis of
the Cesar-Rancheria Basin, Colombia: Implications for it’s tectonic history and
petroleum systems, in C. Bartolini and P. Mann, eds., Petroleum geology and
potential of the Colombian Caribbean Margin: AAPG Memoir 108, p. 431–470.
TOC

Integrated Structural and Basinal Analysis


of the Cesar–Rancheria Basin, Colombia:
Implications for its Tectonic History and
Petroleum Systems
Javier Sanchez and Paul Mann
Department of Earth and Atmospheric Sciences, University of Houston, 312 Science & Research 1,
Houston, Texas 77204, U.S.A.
(e-mails: [email protected]; [email protected])

Abstract
The complex tectonic evolution of northwestern South America is recorded by a ­variety of
­deformed, onland basins ranging in age from Paleozoic to recent. We integrate the M ­ esozoic to
­recent structural geology and basinal history of the 12,000 km2, intermontane C
­ esar-Rancheria
Basin (CRB) in northern Colombia to reconstruct the tectonic history of the basin and tectonic
controls on its sedimentation, subsidence history, and petroleum systems. Methods and data
used for this study include (1) interpretations of crustal structure using gravity and m ­ agnetic
modeling; (2) subsurface mapping of key horizons using 3500 km (2175 mi) of ­two-dimensional
(2-D) seismic reflection data tied to 16 wells; (3) construction of serial structural cross sections,
which are balanced to the level of Late Cretaceous sedimentary rocks; and (4) construction of
burial history graphs and development of a 2-D basin model showing predicted oil and gas
windows. Basement-involved southeast-dipping reverse faults expose Jurassic to early Cre-
taceous rocks in the CRB and Paleozoic rocks in the Perija range (PR) east of the CRB. About
10% of shortening across the CRB is a consequence of motion along these faults driven by
discrete interactions between the Caribbean and South America plates. Two major periods
of NW-SE crustal, interplate shortening are identified: (1) an early–middle Eocene, west-to-
east shortening event that produced east-dipping Cretaceous and Paleocene strata beneath a
­major unconformity that increases in erosional hiatus from east to west across the study area;
this shortening and overlying unconformity is related to the collision of the Great Arc of the
­Caribbean with the northwestern continental margin of the South ­American plate; and (2) a
late Miocene–Pliocene, west-to-east shortening event with major ­exhumation of the eastern
CRB where faulting of recent sedimentary deposits are related to the ­collision between the
­Panama Arc and northwestern South America. The west-to-east migration of d ­ eformation
during the Cenozoic controlled by the west-to-east, diachronous collision ­between the Great
Arc of the Caribbean and northern South America also led the present-day distribution of pre-
served ­depocenters with a major thickness of the Paleocene-early Eocene section to the east of

Copyright ©2015 by The American Association of Petroleum Geologists.


DOI:10.1306/13531945M1083648

431

13880_ch16_ptg01_431-470.indd 431 10/21/15 11:25 AM


432 Sanchez and Mann

the study area, and a major Miocene depocenter to of the tectonic events on source rock maturation, traps,
the west. and petroleum systems in the CRB.

TOC
Introduction Tectonic and Geologic Setting

The continental margin of northwestern South ­America Located 200 km (124 mi) southeast of the thrust front
has experienced a series of complex deformational of the Caribbean–South American subduction plate
events that have included: (1) Mesozoic continen- ­margin, the CRB is a 12,000-km2 intermontane basin
tal rifting either related to the breakup of Pangea and (Figure 1B) bounded by the Sierra de Santa Marta massif
­separation of North and South America or back-arc (SSMM) to the northwest and the Perija range (PR) to the
opening related to subduction along the Pacific Mar- southeast. The SSMM is a remarkably high, (ca 5700 m
gin of South America (Maze, 1984; ­Zuluaga et al., 2015); [18,700 ft]) mountain range with a pronounced trian-
(2) Mesozoic and Cenozoic accretion of the C ­ aribbean gular morphology in map view that reflects the bound-
Oceanic Plateau, the Great Arc of the ­Caribbean, and ing strike-slip faults of the triangular Maracaibo Block
other oceanic and continental fragments by thrust or (Figure 1B). This area overlies the subducting Carib-
strike-slip faulting (Mann et al., 2006; Kennan and Pin- bean oceanic and oceanic plateau crust to the southeast
dell, 2009; Bayona et al., 2011; Cardona et al., 2011a; beneath northern Colombia (Bernal-Olaya et al., 2015a).
Escalona and Mann, 2011); (3) Cenozoic subduction
­polarity reversal that has culminated in the present-day
subduction margin of northern Colombia (­ Kroehler et al., Sierra de Santa Marta massif
2011; Bernal-Olaya et al., 2015b); and (4) Late Miocene
collision of the Panama Arc with the margin of north- The SSMM—a metamorphic massif that bounds
western South America with consequent, northward the CRB along its western edge—includes three
tectonic escape of the triangular Maracaibo Block (Mann lithological belts of exhumed, deeper crustal rocks
and Burke, 1984; ­Vargas and Mann, 2013) (Figure 1). (Figure 2): (1) the oldest belt in the southeast is a
These tectonic events have been recorded in the Meso-Proterozoic (1.2–1 Ga) assemblage of high-
structure and sedimentary history of basins within grade metamorphic rocks that include granulite,
the continental margin area of northwestern South anorthosite, and gneiss that are unconformably
America. In this study, we document the Mesozoic overlain in some areas by Mesozoic and P ­ aleozoic
and Cenozoic sedimentary section and deformation sedimentary and Jurassic volcaniclastic rocks;
history of the Cesar–Rancheria Basin (CRB) located (2) a central belt includes Paleozoic amphibolite, mica
near the northwestern apex of the triangular Marac- schist, and Permo-Carboniferous mylonitic granitoid
aibo Block of northwestern South America (Figure 1A rocks; and (3) a northwestern belt includes Cretaceous
and B). The objective of this chapter is to improve our schist, amphibolite, meta-volcanics, and orthogneiss
understanding of: (1) crustal structure underlying the (Tschanz et al., 1974; Cordani et al., 2005; Cardona
CRB including the locations of inherited M ­ esozoic ­M olina, 2006; Cardona et al., 2010) (Figure 2). All
features; (2) the timing, degree of shortening, and three belts are pervasively intruded by plutonic bod-
driving tectonic mechanisms of Late Cretaceous– ies that include Jurassic-Early Cretaceous and Paleo-
Cenozoic events; (3) the role of the regional strike-slip gene undeformed granites in the southeastern area;
boundaries of the triangular Maracaibo Block on the Paleozoic and Paleogene syn-tectonic granitoid intru-
­development of its overlying basins including the CRB sive rocks in the central area; and Late Cretaceous
and the Maracaibo Basin to the east; and (4) the effects and Paleocene undeformed intrusive rocks in the

Figure 1. Location map showing the regional tectonic and geologic setting of the Caribbean-South American plate boundary
in northwestern South America. (A) The main plate boundaries in the Caribbean region with GPS vectors in red show the
motion of the Caribbean plate relative to the South American plate (modified from Calais and Mann, 2009). The red box
shows the area of the residual gravity map in B. (B) Residual gravity map showing major crustal and tectonic features of
northwestern South America. The gray line shows the location of the gravity and magnetic profile in the Figure 4A and the red
box corresponds to the map area shown in Figure 2. CF 5 Cuisa fault, CP 5 Caribbean plate, CRB 5 Cesar-Rancheria Basin,
GB 5 Guajira Basin, LMV 5 Lower Magdalena Valley Basin, MB 5 Maracaibo Basin, MD 5 Magdalena delta, OF 5 Oca fault,
PR 5 Perija Range, RF 5 Romeral suture, SBF 5 Santa Marta-Bucaramanga fault, SCDB 5 South Caribbean deformed belt,
SM 5 possible seamount, SSMM 5 Sierra de Santa Marta massif, TF 5 El Tigre fault. 100 km (62.1 mi)

13880_ch16_ptg01_431-470.indd 432 10/21/15 11:25 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 433

TOC

13880_ch16_ptg01_431-470.indd 433 10/21/15 11:25 AM


434 Sanchez and Mann

northwestern area (Tschanz et al., 1974; Cardona et al., Ayala, 2009) including the Cerrejon thrust fault along its
2010; Villagomez et al., 2011). The southeastern and northern range front (Figure 2). The eastern PR is bounded
central belts are separated by the Sevilla lineament that by northwest-dipping thrusts including the Motilones
was interpreted by previous workers as a high-angle, thrust fault along the southern edge of the range (Figure 2) TOC
­southeastwardly-dipping thrust fault. The central belt and triangular zones where high-angle basement-involved
overthrusts the northwestern belts along the high-­ thrusts accommodate displacement by backthrusting
angle Aguja reverse fault that also dips to the s­ outheast along the Cretaceous detachment section (Duerto et al.,
(Tschanz, 1974; ­Villagomez et al., 2011) (­Figure 2). 2006). Convergent deformation has been accommodated
by thrusting in the PR with a total thin-skinned shorten-
ing of 46–56 km (28.6–34.8 mi) proposed by Kellogg and
Perija Range Bonini (1982). Other workers have proposed a thick-
skinned structural style and triangle zone geometry along
The NE–SW-trending PR forms the ­eastern margin of the the eastern margin of the PR with 4–6 km (2.5–3.7 mi) of
CRB and separates the CRB from the western Maracaibo shortening (Gallango et al., 2002; Duerto et al., 2006).
Basin in Venezuela (Figure 1). The oldest rocks of the PR Early Eocene (ca 53 Ma), middle Eocene (ca 45
include Cambro-Ordovician phyllite, schist, quartzite, Ma), and late Oligocene (ca 25 Ma) unconformities
and unconformable meta-conglomerate that are in turn have all been reported along the eastern flank of the
overlain by Devonian to Permian sandy and calcare- PR ­(Kellogg, 1984). These unconformities were inter-
ous marine rocks that are in turn overlain by Jurassic, preted by Kellogg (1984) as evidence for the onset of
arkosic redbeds inter-layered with volcanoclastic and convergent deformation and uplift of the PR by early
volcanic rocks (Maze, 1984) (Figure 3). Early Cretaceous Eocene with an increase in the uplift rate and thrust-
conglomerate and arkosic sandstone unconformably ing during the late Oligocene. Thrusting produced
overlies Jurassic rocks and are conformably overlain 3 to 4 km (1.9 to 2.5 mi) of topographic elevation of
by middle Cretaceous limestones and black calcareous the PR and unroofing of the thrust belt (Kellogg, 1984).
shale (Caceres et al., 1980). Paleogene shale, calcareous Thickness changes of upper Paleogene strata exposed
shale, limestone, and sandstone conformably over- along the eastern flank of the PR also suggest that con-
lie Cretaceous rocks in the northern PR (Forero, 1970; vergence along the Motilones thrust fault (Figure 2)
­Caceres et al., 1980; Ujueta and Llinas, 1990) (Figure 3). occurred during the Oligocene (Duerto et al., 2006).
The El Tigre–Perija fault forms a prominent line-
ament that parallels the length of the PR (Figure 2).
The sense of displacement of the El Tigre-Perija fault Cesar-Rancheria Basin
has been determined to be either left-lateral (Duerto
et al., 2006; Ayala, 2009) or right-lateral (Miller, 1962) The CRB is filled by a 2-km-thick (1.2-mi) clastic wedge
with a significant right-lateral displacement 15–35 km eroded from the southeastern flank of the SSMM
(9.3–21.7 mi) (Alberding, 1957) and a later Eocene– that includes pre-­Cretaceous continental, Cretaceous
Cretaceous normal reactivation with the eastern block marine, and C­ enozoic continental sedimentary sections
of the fault zone upthrown by 1500 m (4921 ft) along (Bayona et al., 2007; Ayala, 2009). Neighboring basins
the Totumo-Inciarte arch (Miller, 1962) (Figure 2). of the CRB include the Lower Magdalena Valley (Plato
The El Tigre fault has also been interpreted as a 10° Subbasin) to the southwest and Baja Guajira Basin
to 20° southeastward-dipping thrust fault that accom- to the north (Figure 1B) with total sedimentary thick-
modated northwest-southeast shortening across the nesses ranging from 8 km (4.9 mi) (Duque-Caro, 1979)
range that began in the Eocene (Kellogg, 1984). and 3 km (1.9 mi) (Rincon et al., 2007), respectively.
The western flank of the PR is characterized by south- The CRB is subdivided into two subbasins: the
eastward-dipping thrusts (Kellogg and Bonini, 1982; southern Cesar (CB) and the northern Rancheria (RB)

Figure 2. Generalized geologic map of the Sierra de Santa Marta massif (SSMM) and Perija range (PR) for which colors show
rock age and patterns show lithologies (see map key for explanations). Important structural features include the Aguja fault
(AF), Sevilla lineament (SL), Cerrejon fault (CF), El Tigre fault (TF), Los Motilones fault (MF), and Totumo arch (TA). The
Cesar–Rancheria Basin (CRB) is divided into the Cesar (CB) and Rancheria (RB) Subbasins. Wells are shown as gray dots and
two-dimensional (2-D) seismic lines are shown as blue lines. Outcrops of Jurassic and Cretaceous rocks are exposed along
the Verdesia High (VH) in the center of the CRB. Adjacent basins include: (1) Lower Magdalena Valley (LMV) to the southwest
separated from the CRB by the Santa Marta–Bucaramanga left-lateral, strike-slip fault (SBF); (2) the Maracaibo Basin (MB)
to the east; and (3) Guajira Basin (GB) to the north separated from the CRB by the Oca right-lateral, strike-slip fault (OF).
Numbered circles show the locations of different burial history graphs (Figure 10). GPS vector shows rate of motion of the
northwestern South America plate relative to a fixed Caribbean plate. 20 km (12.4 mi)

13880_ch16_ptg01_431-470.indd 434 10/21/15 11:25 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 435

TOC

13880_ch16_ptg01_431-470.indd 435 10/21/15 11:25 AM


436 Sanchez and Mann

Subbasins (Figures 1 and 2) that are separated in the The western Maracaibo Basin of Venezuela contains
center of the basin by the Valledupar basement high correlative Mesozoic units to the CRB including the
(Figure 8B and E). The CB includes the Verdesia high, the Jurassic la Quinta Formation cropping out along
a southern continuation of the Valledupar High that its basin margins (Maze, 1984; Lugo and Mann, 1995). TOC
is a large, fault-controlled anticline exposing J­ urassic The Middle to upper Cretaceous interval includes car-
rocks in its core (Ayala, 2009). Other important bonate and mudstone facies related to passive margin
NE–SW-trending structures are the Venados syn- deposition (Audemard, 1991; Castillo, 2001; Mann et
cline in the western CRB and the Becerril anticline al., 2006). The Paleocene Guasare and early Eocene
in the e­ astern CRB (Figure 8A and B). The CB is Misoa Formation in the western Maracaibo Basin
­limited to the south by the Santa Marta–Bucaramanga show lateral facies changes to limestone and shallow
fault (SBF) system that is a regional, active, left-lat- marine, respectively (Parnaud et al., 1995; Mann et al.,
eral ­s trike-slip system with a proposed left-lateral 2006; Escalona and Mann, 2011) in contrast to the more
­displacement of 30–110 km (18.6–68.3 mi) (Campbell, continental Cerrejon Formation in the RB ­(Figure 3).
1965; ­F lorez-Nino, 2001). The SBF does not have a ­Thepper Eocene and Oligocene units present in the
­well-defined trace in the boundary area between the western Maracaibo include non-marine predominantly
Lower Magdalena Valley Basin and the CRB (Ujueta, sandy facies (Sierra, Ceibote, and Icotea Formations),
2003; Mora and Garcia, 2006) (Figures 1 and 2). The while in the CRB this sedimentary record is absent.
SBF has been interpreted as a Mesozoic right-lateral The M­ iocene and Pliocene non-marine section in the
fault that controlled the formation of several Meszoic western Maracaibo Basin (Macoa, Los Ranchos, and El
rift basins to the south (Kammer and Sanchez, 2006). Milagro Formations) is correlative with similar facies
The RB is an elongated depression filled mainly by a in the CRB (Parnaud et al., 1995; Mann et al., 2006).
Mesozoic and Cenozoic clastic wedge with interbedded
carbonate rocks deposited on the southeastern flank of
the SSMM and later buried by ­Quaternary alluvium Cenozoic stratigraphy of the CRB
(Figure 3). Large structures such as the Tabaco anticline
expose Eocene and Paleocene rocks to the north of the The Cenozoic sedimentary wedge in the CRB indicates
subbasin (Caceres et al., 1980; Montes et al., 2010) (­Figure tilting and unroofing of the SSMM by the Paleocene
8A). The southeastward-dipping Cerrejon fault (Figure (Bayona et al., 2007) associated with the Caribbean–
2) forms the eastern boundary between the RB and the South American plate collision (Mann et al., 2006) and
PR (Miller, 1962; Kellogg and Bonini, 1982; ­Kellogg, accretion of a P­ aleocene Arc (Cardona et al., 2011a,
1984; Montes et al., 2010). It has been interpreted as a b). This collision would have tilted the northwest-
16- to 26-km (9.9- to 16.1-mi) displacement low-dipping ern ­corner of the SSMM and created accommodation
thrust fault with a detachment at a depth of 8 km (4.9 space along its southeastern flank, where accumula-
mi) (Kellogg and Bonini, 1982; Kellogg, 1984). This tion of the C
­ enozoic units took place (Montes et al.,
subbasin is bound to the north by the Oca fault (Fig- 2005b) in a general southeast-dipping monocline that
ure 2), a right-­lateral strike-slip system with 15–20 km maintains its structural continuity with the SSMM
(9.3–12.4 mi) (Feo-­Codecido et al., 1984) to 50–90 km (Figures 2 and 4B).
(31.1–55.9 mi) (Kellogg, 1984) of post-Eocene displace- Cenozoic sedimentation mainly consisted of con-
ment. Both the Santa Marta and Oca fault activity tinental deposits with some shallow-water carbon-
started by the ­Oligocene (Kellogg, 1984; Villamil, 1999). ate rocks found in the northern CRB. The sandy
Barco ­Formation of Paleocene age transitions into the
muddy Cuervos Formation in the southern CRB. In
Cretaceous stratigraphy of the CRB the northern CRB, shallow marine calcareous rocks
of the Paleocene Hato Nuevo and Manantial Forma-
The Cretaceous section of the CRB was deposited in tions transition into a continental mudstone and coal-
a passive margin setting generally consisting of shal- bearing unit (Cerrejon Formation) (Bayona et al., 2007;
low marine inner-shelf carbonate rocks and outer- Ayala, 2009). (Figure 3).
shelf mudstone (El Molino, Socuy, Tres Esquinas, and Two major unconformities in the middle Eocene–
La Luna ­Formations, Cogollo Group). Lower Creta- Oligocene and in the upper Miocene–Pliocene are pre-
ceous sedimentation of continental sandstone and sent in the CRB. Predominantly sandy units with some
conglomerate (Rio Negro Formation) took place in interbedded continental mudstone and conglomerate
an extensional rift setting along with the underlying were deposited during the Miocene to the south (Cuesta
unconformable Jurassic rocks (La Quinta Formation) Formation) and are correlative to calcareous and con-
that contain sandy and conglomeratic redbeds and glomeratic units to the north (lower ­Miocene Conjunto
tuffs, andesite flows, and ash deposits (Caceres et al., Calcareo and upper Miocene C ­ onjunto Conglomertico
1980; Mann et al., 2006) (Figure 3). Formations, (Caceres et al., 1980). ­Conglomeratic and

13880_ch16_ptg01_431-470.indd 436 10/21/15 11:25 AM


PERIOD EPOCH Ma LEGEND
CR (2) Perija range (3) Maracaibo (4)
El Milagro El Milagro
Pliocene Gallinas La Puerta Non-marine conglomerate
Castilletes
10 Cuesta La Villa La Villa Non-marine mudstone and siltstone
Jimol Los Ranchos Los Ranchos

13880_ch16_ptg01_431-470.indd 437
Miocene Cuiba Cuiba
Macoa Macoa Non-marine sandstone

Neogene
20 Utipa
Icotea
Siamana Peroc Coastal plain sandstone
Oligocene 30 Ceibote
Coastal plain and lacustrine mudstone
Nazareth Sierra
40 Shallow-marine sandstone
Macarao
Eocene Mirador (La Loma) Mirador Shallow-marine inner shelf mudstone
50 Misoa
Outer-shelf shale and carbonate

Paleogene
Mirador
Cuervos Marcelina
Paleocene 60 Orocue Chert
Barco Guasare Guasare
Ma El Molino Shallow-marine inner shelf carbonate
70 Mito Juan
Socuy Colon
Felsic tuff
Socuy
Ca 80 Tres esquinas Tres esquinas Intermediate to acidic intrusive rocks
GAC

Late
Sa
Co La Luna Coal
90 La Luna
Tu
Ce Ash
100 Aguas Blancas
Maracas Source rock
Al
110 Lisure

Cretaceous
Ap Lagunitas
120 Apon
1

Early
Ba 130 Rio Negro
He
Va 140
Be 3
150 La Quinta La Quinta
Late
160 2
Middle 170 Macoita Macoita

Jurassic
4
Tinacoa
180 Tinacoa
Early
NW CF
SE
25 km SSMM TF

Figure 3. General stratigraphic chart of northwestern South America showing the main sedimentary formations for the offshore area of the adjacent Caribbean Sea,
the Cesar Rancheria Basin (CR), the Perija range, and the Maracaibo Basin. Inset map shows the location of these four areas. Stratigraphic data were compiled from
Miller (1962), Forero (1970), Caceres et al. (1980), Kellogg (1984), Maze (1984), Parnaud et al. (1995), Mann et al. (2006), Vence (2008), and Ayala (2009).
SSMM 5 Sierra de Santa Marta massif; CF 5 Cerrejon fault; TF 5 El Tigre fault. 25 km (15.5 mi)
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 437

TOC

10/21/15 11:26 AM
438 Sanchez and Mann

sandy Pleistocene rocks (El Milagro Formation) discor- for Cretaceous metamorphic and Paleocene intrusive
dantly overlie Miocene units (Figure 3). rocks occur near the northwest part of the CRB with
the youngest exhumation pulse occurring by the mid-
dle Miocene (ca 16 Ma). The high, modern topographic TOC
Links between the CRB and relief of the SSMM may be the result of a very recent
neighboring basins surface uplift (<1 Ma) where the rocks that record this
event have not yet been exposed (­ Villagomez et al.,
­ revious studies propose that the Middle Magdalena
P 2011). U–Th/He thermochronology in apatites and
Valley (MMV) Basin and CRB were once continuous zircons from Eocene granitoid intrusions show that
basins that became tectonically separated during the Miocene uplift (ca 25–5 Ma) occurred in two pulses
Cenozoic. Four mechanisms have been proposed to of elevated exhumation from the early to middle
­explain this tectonic block separation: (1) vertical-axis ­Miocene (ca 24-15 Ma) (Cardona et al., 2011b).
rotation of the SSMM Block (ca 34°) (Duque-Caro, 1979; For Cretaceous rocks of the Cogollo Group, ther-
Reyes-Santos et al., 2000; Montes et al., 2010); (2) 110 km mochronology based on AFT indicate that the most
(68.3 mi) of left-lateral displacement along the SBF recent exhumation event occurred by the late Miocene
(Campbell, 1965); (3) in situ northwest to southeast con- (ca 10–5 Ma) (Figure 4C). The thermal history of this
vergence, deformation, and uplift of the SSMM and PR area prior to this event has been reset by total annealing
(Villamil, 1999; Cediel et al., 2003); or (4) the vertical-axis (Hernandez and Jaramillo, 2009). ­Apatite fission tracks
rotation (50°–75°) of the entire MB (Montes et al., 2005a). in northeastern PR indicates an uplift pulse (Figure 4C)
by late Oligocene–early ­Miocene (22–27 Ma) for the
Jurassic Palmar granite and Paleozoic meta-­sedimentary
Previous provenance and exhumation studies rocks in the Palmar High (Figure 2) and a Pliocene (ca 3
Ma) uplift pulse (Figure 4C) in the Totumo–Inciarte arch
Sandstone petrography, provenance, paleocurrent, (Figure 2) that affected the Jurassic La Quinta Formation
and sedimentological analyses in the Paleocene and Paleozoic Lajas granite (Shagam et al., 1984).
Manantial and Cerrejon Formations in the northern-
most part of the CRB indicate that the clastic sedimen-
tary section (>1500 m [4921 ft] thick) was derived from Methods
the erosion of the SSMM along the western flank of
the CRB (Bayona et al., 2007; Ayala, 2009). This syn- Methods used in this study include: (1) seismic inter-
orogenic sedimentary wedge thickens to the south- pretation and sub-surface mapping of 2-D seismic
east with a maximum thickness of 2.2 km (1.4 mi). profiles, (2) potential field (gravity and magnetics)
Increasing ­K-feldspar and quartz content near the con- analysis and modeling, (3) one- and two-dimensional
tact with the overlying lower Eocene Tabaco Forma- basin modeling, and (4) kinematic analysis by quanti-
tion indicates a PR provenance from the PR followed tative restoration of regional, structural cross sections.
by Eocene uplift (Bayona et al., 2007; Ayala, 2009;
­Cardona et al., 2011a).
Thermochronological analysis using apatite fission Potential Fields Analysis
tracks (AFT) has revealed the onset of exhumation for
Jurassic intrusive rocks in the southeastern SSMM by The gravity dataset used in this study include Bouguer
the Paleocene (ca 65–58 Ma). Higher exhumation rates anomaly data kindly provided by the Agencia Nacional

Figure 4. A) Regional gravity and magnetic modeling used to infer the crustal structure beneath the Sierra de Santa Marta
massif (SSMM) and the study area of the CRB. These two, topographically elevated, onland features occupy the northern
apex of the fault-bounded Maracaibo Block shown in map view on the inset map and overlie the southeastward-subducting
slab of the Caribbean plate. The dashed, black box shown on the cross section in A represents the position of the regional,
geologic cross section shown in B. B) This geologic cross section includes the SSMM, Cesar–Rancheria Basin (CRB), Perija
Range (PR), and Maracaibo Basin (MB), where the Tigre fault (TF) in the PR is a major regional fault. L­ ocations of previously
published thermochronologic (orange dots) and geochronologic (red dots) dates are plotted on the cross section. C) Age
vs. distance plot for thermochronologic ages shown on the cross section in B. D) Age vs. distance plot for geochronologic
ages shown on the cross section in B. Thermochronologic and geochronologic ages plotted in C and D were compiled from
previous studies by: Tschanz et al. (1974), Shagam et al. (1984), MacDonald and Opdike (1984), Restrepo-Pace et al. (1997),
Cordani (2005), Cardona Molina et al. (2006), Hernandez and Jaramillo (2009), Bayona et al. (2010), Cardona et al. (2010,
2011a, b), and Villagomez et al. (2011). See text for discussion.

13880_ch16_ptg01_431-470.indd 438 10/21/15 11:26 AM


NW SE Water: Middle continental crust:
A ρ=1.03 g/cm3 ρ=2.9 g/cm3
150 mag. sus.= 0 u-cgi mag. sus.= 8000 u-cgi
100 Upper Cenozoic: Lower continental crust:
50 ρ=2.2 g/cm3 ρ=3.06 g/cm3
0 mag. sus.= 0 u-cgi mag. sus.= 16000 u-cgi

-50 Lower Cenozoic: Grenvillian basement


ρ=2.4 g/cm3 (continental crust):

13880_ch16_ptg01_431-470.indd 439
-100 mag. sus.= 0 u-cgi ρ=3.06 g/cm3

Magnetics (nT)
-150 Cretaceous-Pre-Cretaceous: mag. sus.= 4000 u-cgi
-200 =Observed, =Calculated, =Error 18.472 ρ=2.5 g/cm3 Oceanic crust:
mag. sus.= 0 u-cgi ρ=2.9g/cm3
200 mag. sus.= 15000 u-cgi
Acidic intrusive bodies:
150 ρ=2.67 g/cm3 Upper mantle crust:
mag. sus.= 4000 u-cgi ρ=3.36 g/cm3
100 mag. sus.= 68 u-cgi
Upper continental crust:
50 ρ=2.67 g/cm3
Refraction data control
mag. sus.= 15000 u-cgi
0 =Observed, =Calculated, =Error 8.443

Gravity (mGal)
-50
-100
A A’
0 B B’
lower Cenozoic
10 upper crust
20 moho

30

Depth (km)
40 20 km
VE =1
50

B SSMM CRB PR MB B’
0
4
8 TF
12 5 km
km

C 60

40

20 A
Age (Ma)

0
10000 20000 30000 40000 50000 60000 70000 80000 90000 100000 110000 120000 130000 140000 150000 160000 170000 180000 190000 200000 210000 220000 230000 240000 250000 260000

D 2000
1500
1000
500
Age (Ma)

A’
0
10000 20000 30000 40000 50000 60000 70000 80000 90000 100000 110000 120000 130000 140000 150000 160000 170000 180000 190000 200000 210000 220000 230000 240000 250000 260000

Sierra de Santa Marta Massif Cesar-Rancheria and Perija


Santa Marta Santa Marta Plio/Pleistocene Early Cretaceous Thermochronology Geochronology
metamorphic belt batholith
Late Miocene Jurassic/Triassic AFT U-Pb
Metadiorite
Sevilla Early Miocene Rb / Sr
metamorphic belt Paleozoic U -Th /He Zircon
Central batholith Oligocene/Eocene K/ Ar , Ar / Ar
Los Mangos Paleocene Basement U -Th /He Apatite Depositional age
granulite Pueblo Bello and
Late Cretaceous
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 439

Patillal batholith
TOC

10/21/15 11:26 AM
440 Sanchez and Mann

de Hidrocarburos (ANH) of Colombia, which was cal- crustal sources of the regional gravity anomaly (Jacob-
culated using a Bouguer correction of 2.7 g/cm3. This sen, 1987; Zeng et al., 2007); by calculating the residu-
Bouguer gravity dataset was based on a merge of the als, we enhance the shallow-source and local gravity
following surveys: (1) a detailed aero-gravimetry grid anomaly; (2) reduction to the pole (RTP) of the total TOC
(ca 1 km [0.6 mi] sample s­ pacing between NE–SW magnetic field, which calculates the magnetic anomaly
transects and 4.5 km (2.7 mi) between the transects); that would produce a source if the anomaly were at
these data cover the Rancheria Subbasin (RB) (Figure the geomagnetic pole and was magnetized vertically
2); (2) land gravimetry (1–12 km [0.6–7.5 mi] of sample (Baranov and Naudy, 1964); RTP produces more sym-
spacing) that covers the Cesar Subbasin (CB); and (3) metrical anomalies above the magnetic sources and
our addition of digitized points from the downloaded allows a more precise correlation with geologic fea-
ANH total Bouguer anomaly map for areas with sparse tures; and (3) tilt derivative filters were applied to both
data coverage or no data coverage—especially in the the Bouguer anomaly and RTP magnetic grids; these
area of the SSMM and PR (Figure 2). derivative filters enhance the gradient of the signal and
We used a satellite-gravity grid (Sandwell and help identify edges and the lateral continuity of crustal
Smith, 1997) of free-air anomaly (ca 8 km [4.9 mi] sam- structures (Nettleton, 1971; Verduzco et al., 2004).
ple spacing) to construct a regional 2-D forward model Our regional 2-D forward model shown in Figure 4A
including an offshore area to the west of the SSMM was constrained by: (1) refraction data that constrains
(Figure 4A). Bouguer anomaly offshore was obtained the depth to the Moho of the offshore Caribbean Oce-
by adding a calculated Bouguer correction (using 0.9 anic Plateau at 16–20 km (9.9–12.4 mi) (Ewing et al.,
g/cm3) to the free-air anomaly. 1960) (Figure 1); (2) deep-penetration seismic reflec-
The magnetic dataset used in our analysis was also tion lines in depth that show the top of the subduct-
provided by ANH and consists of a detailed aero-mag- ing Caribbean Oceanic Plateau at a depth of 10–13 km
netic grid (ca 1 km [0.6 mi] sample spacing) of the total (6.2–8.1 mi) (Bernal-Olaya et al., 2015b); (3) tomographic
magnetic field covering the areas of the CRB, part of the imagery of the subducting Caribbean plate beneath
SSMM, and western PR (Figure 2). To create a regional South America that shows that the Caribbean Oceanic
forward model (Figure 4A) that included the Carib- Plateau dipping to the southeast at an angle of 15°–20°
bean Sea west of the SSMM, we used the global earth (van der Hilst and Mann, 1994); (4) tabular alignment
magnetic anomaly grid (EMAG2) that is a merger of of earthquakes to depths of 300 km (186.4 mi) that are
satellite, ship, and airborne measurements (Maus et al., interpreted as the Benioff zone produced by southeast-
2009). Enhanced gravity and magnetic grids were cal- ward subduction of the Caribbean Oceanic Plateau
culated from the Bouguer anomaly and total magnetic (Kellogg and Bonini, 1982; Bernal-Olaya et al., 2015a);
field grids, respectively, for the purpose of identifying (5) geologic map information for the SSMM, CRB, and
structural and tectonic features within the crystalline PR taken from previous workers (Gomez Tapias et al.,
basement using Geosoft Oasis Montaj software v. 7.3. 2007); and (6) our seismic interpretation for the base of
These enhanced gravity and magnetic grids were the Cretaceous and the Cenozoic sections in the CRB. In
created using the following methods: (1) a Bouguer addition, six local gravity and magnetic models were
residual map was made by subtracting the Bouguer constructed along the CRB to define the top of basement
anomaly grid from an upward continuation filter at and overall configuration of the CRB (Figure 5A and D).
10 km (6.2 mi); continuation methods treat the signal Results from the regional 2-D model were used to
amplitude changes above the source; in the case of the estimate the depth to the Moho and to the Caribbean
upward continuation filter, this method emphasizes subducting slab beneath northern Colombia. Forward
long-wavelength anomalies that enhance the deeper modeling was performed using the GM-SYS (Geosoft

Figure 5. Potential field maps used to illustrate regional tectonic and structural features of the study area: (A) Bouguer
anomaly map displaying main tectonic provinces and showing the outline (white polygon) of the Cesar–Rancheria Basin
(CRB) and locations of the gravity and magnetic modeling shown in Figure 6; (B) residual Bouguer anomaly after an upward-
continuation filter showing important structural features explained in the caption of Figure 2; (C) tilt derivative calculation of
the Bouguer anomaly identifying possible faults; (D) total magnetic field showing main tectonic provinces, the outline of the
CRB, and locations of the sections used for gravity and magnetic modeling in Figure 6; (E) reduction to the pole (RTP) map
for the total magnetic fields that correct the positions of anomalies and improve identification of differing basement blocks;
(F) tilt derivative map of the total magnetic field used to define linear faults possibly within the crystalline basement.
CB 5 Cesar Subbasin, CF 5 Cerrejo fault, CL 5 Cesar lineament, OF 5 Oca fault, PR 5 Perija range, RB 5 Rancheria Subbasin,
SBF 5 Santa Marta–Bucaramanga fault, SL 5 Sevilla lineament, SSMM 5 Sierra de Santa Marta massif, VAH 5 Valledupar
High, VH 5 Verdesia high. 10 km (6.2 mi)

13880_ch16_ptg01_431-470.indd 440 10/21/15 11:26 AM


A. Bouguer anomaly B. Residual after upward continuation (10 km) C. Tilt derivative

Se
ct
io
n
4

13880_ch16_ptg01_431-470.indd 441
OF OF

Se
ct
SSMM

io
RB CL CL

n
3
CF CF

Se
c tio
n
2

Se
c
VAH VAH

tio
n
1
PR
VH VH
SBF SBF
CB
n5
c tio
Se
6
n
c tio
Se

Rad

E. RTP F. Tilt derivative

Se
ct
io
n
4

OF OF

Se
c
SSMM RB SL SL

tio
n
3
CF

Se
ct
io
n
2

Se
c
tio
n
1
PR VAH VAH

VH VH
CB SBF
5 SBF
i on
ct
Se
n6
io
ct
Se
Rad
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 441

TOC

10/21/15 11:26 AM
442 Sanchez and Mann

software, v. 7.3). For input data, we integrated the integration of information (unit thicknesses, litholo-
Bouguer anomaly grid for the onshore area, the free gies, ages, and current bottomhole temperature) from
air grid for the offshore area, and the total magnetic seven wells; a published stratigraphic column (Bayona
field grid. et al., 2007); a pseudo-well created in the center of the TOC
We created models using: (1) a water density of basin using sub-surface mapping from seismic inter-
1.03 g/cm3; (2) three sedimentary layers with densities pretation (Figure 2); and a 2-D basin model based on a
ranging from 2.4 g/cm3 to 2.6 g/cm3; (3) an upper crus- NW–SE structural cross section across the southern
tal layer of 2.7 g/cm3 density and 0.05 SI magnetic sus- part of the basin (Figure 2). The burial history dia-
ceptibility; (4) a lower crustal layer and the Caribbean grams were constructed and simulated in the Petro-
Plateau of 2.9-g/cm3 density and a 0.1 SI magnetic sus- mod software v. 2012. The bottomhole temperature
ceptibility; (5) metamorphic rocks with density of 2.7 g/ measurements were used to calibrate the recent heat
cm3 and magnetic susceptibility of 0.005 SI; and (6) an flow of the CRB, which is 40 mW/m2.
upper mantle layer of 3.3 g/cm3 density and 0.008 SI We used the McKenzie (1978) model to estimate
magnetic susceptibility (Figure 6). The modeling pro- the heat flow for the Mesozoic that included a long-
cess consists of obtaining valid crustal and basin geom- lived period of Jurassic rifting (Maze et al., 1984)
etries by comparing the synthetic signal generated by that reached its highest heat flow of 70 mW/m2 in
the model with the observed gravity and magnetic data the Jurassic and decreased to the Cretaceous passive
along the profiles and seeking to find the best match margin phase to 60–45 mW/m2. During the Cenozoic
between both curves (Blakely, 1996) (Figure 6). characterized by periodic convergent events, heatflow
further decreased to 40–45 mW/m2.
Paleo-water depth was estimated by using depo-
Seismic Interpretation and Mapping sitional environments for the different units where
deeper facies (outer neritic) are Cenomanian–Turonian
Sub-surface interpretation in this study is based on and shallower to continental by Jurassic and Cenozoic.
mapping of 3100 km (1926 mi) of 2-D seismic data tied Miocene carbonatic facies in the northern RB denotes a
to 16 wells within the CRB (Figure 2). The seismic data shallow platform. The sediment–water interface temper-
and wells were kindly provided by ANH and OGX. ature (SWIT) was automatically assigned in ­Petromod
Seismic interpretation and subsurface mapping was (Wygrala, 1989) by introducing a latitude value.
carried out using Petrel v. 2011 software. Source rock parameters including total organic
Four synthetic seismograms were created using ­c arbon (TOC) and hydrogen index (HI) were com-
sonic, density, and check-shot logs available for seven piled from well reports and published data (Garcia
of the 16 wells used in the study (Figure 2). They were et al., 2008; Ayala, 2009) and used to constrain the
used to tie well tops in depth to 2-D seismic lines in hydrocarbon transformation ratio (TR), which is a
two-way time. Once we established the seismic char- quantification of the amount of hydrocarbons gener-
acter of key lithological units and reflectors, we made ated by chemical reactions in source rocks calculated
a regional interpretation and mapping in two-way by a kinetic model (Tissot and Welte, 1984). The TOC
time (TWT) for five horizons that included mapping values are 1–3% for the Molino Formation, 1–5% for
of fault polygons for each horizon (Figure 8). Surfi- the La Luna Formation, and 1–4% for the Cogollo
cial geologic maps and structural measurements by Group. The HI values are 100–400 mgHc/g TOC for
previous workers (Gomez Tapias et al., 2007; Bayona the Molino Formation, 120–600 mgHc/g TOC for the
et al., 2010, 2011) were used in areas of outcrops to La Luna Formation, and 150–500 mgHc/g TOC for the
constrain subsurface units. Structural maps in depth Cogollo Group (Figure 3).
were c­ reated by multiplying interval velocity maps Maximum burial depth is constrained by vitrinite
obtained from check-shot logs and structural maps in reflectance and Tmax data (Tissot et al., 1987) from some
time. To better understand depocenter geometries and wells. The timing of exhumation events has been cali-
the extent of major unconformities, we created four brated with thermochronological studies ­(Figure 4C)
sediment thickness maps by subtracting two vertically along SSMM, CR, and PR (Shagam et al., 1984;
consecutive structural maps. ­Hernandez and Jaramillo, 2009; Cardona et al., 2011a,
b; Villagomez et al., 2011). In general, these exhumation
periods are related to stratigraphic unconformities.
Basin Analysis For the 2-D modeling, we used the southern struc-
tural cross section (Figure 9C), which was edited
Our basin analysis performed on the CRB con- and gridded in the Petromod software. Three burial
sists of nine burial history diagrams constrained by ­history diagrams along the section were considered to

13880_ch16_ptg01_431-470.indd 442 10/21/15 11:26 AM


A. Section 1 B. Section 2 C. Section 3 D. Section 4
NW section 5 section 6 SE NW section 5 section 6 SE NW section 5 section 6 SE NW section 5 section 6 SE
50.00
40.00 -80.00
0.00
0.00
0.00 -160.00
-40.00
-50.00
-40.00 -240.00
-80.00

Magnetics (nT)

13880_ch16_ptg01_431-470.indd 443
Magnetics (nT)
Magnetics (nT)

Magnetics (nT)
=Observed, =Calculated, =Error 4.824 =Observed, =Calculated, =Error 7.712 =Observed, =Calculated, =Error 3.318 =Observed, =Calculated, =Error 6.631
10.00 40.00
80.00
0.00
60.00
-10.00 60.00 20.00

-20.00 30.00 40.00


0.00
-30.00 20.00
0.00

Gravity (mGal)

Gravity (mGal)

Gravity (mGal)

Gravity (mGal)
-40.00 =Observed, =Calculated, =Error 3.162 =Observed, -20.00
=Calculated, =Error 2.313 =Observed, =Calculated, =Error 2.467 =Observed, =Calculated, =Error 1.782

0.00 PR CB PR RB PR RB PR MA
CB 0.00 0.00 0.00
SSMM VH 10.00SSMM ? ? SSMM ? TF TA
20.00 SBF 20.00 20.00 20.00
30.00
40.00 40.00 40.00 40.00

Depth (km)

Depth (km)

Depth (km)
Depth (km)
50.00
0.00 30.00 60.00 90.00 0.00 30.00 60.00 90.00 0.00 20.00 40.00 60.00 0.00 30.00 60.00 90.00
VE =0.5 VE =0.5 VE =0.5 VE =0.5
Scale =1600000 Distance (km) Scale =1600000
Distance (km) Scale =1600000 Distance (km) Scale =1600000 Distance (km)

E. Section 5 F. Section 6
NE section 1 section 2 section 3 section 4 SW NE section 1 section 2 section 3 section 4 SW
0.00

0.00
-30.00

-100.00
-60.00

Magnetics (nT)

Magnetics (nT)
-200.00
-90.00
=Observed, =Calculated, =Error 20.142 =Observed, =Calculated, =Error 5.889
60.00

30.00 30.00

0.00 0.00

-30.00 -30.00

Gravity (mGal)

Gravity (mGal)
=Observed, =Calculated, =Error 2.551 =Observed, =Calculated, =Error 4.96

0.00 PR
0.00 RB CB
SBF CB VH VAH OF ? OF
?

Depth (km)
50.00
Depth (km)
50.00
0.00 80.00 160.00 240.00
VE =0.5 0.00 80.00 160.00 240.00
Scale =1600000 Distance (km) VE =0.5
Scale =1600000 Distance (km)

Cenozoic sedimentary rocks Upper continental crust Upper mantle


(Density: 2400 kg/m3, Susceptibility: 0.0005 SI) (Density: 2670 kg/m3, Susceptibility: 0.05 SI) (Density: 3300 kg/m3, Susceptibility: 0.008 SI)
Cretaceous sedimentary rocks Lower continental crust
(Density: 2450 kg/m3, Susceptibility: 0.0005 SI) (Density: 2900 kg/m3, Susceptibility: 0.1 SI)

Pre-Cretaceous sedimentary and Medium and high grade metamorphic rocks


meta-sedimentary rocks (Density: 2600 kg/m3, (Density: 2800 kg/m3, Susceptibility: 0.005 SI)
Susceptibility: 0.0005 SI)

Figure 6. Gravity and magnetic modeling of crustal structure underlying the region of the CRB based on six modeled sections located on Figure 5A and D.
Parameters used in the modeling are summarized in the legend. The names of the main structural features and their intersections with other sections are
labeled. CB 5 Cesar Subbasin, OF 5 Oca fault, PR 5 Perija range, RB 5 Rancheria Subbasin, SBF 5 Santa Marta-Bucaramanga fault, SSMM 5 Sierra de Santa
Marta massif, TA 5 Totumo arch, TF 5 El Tigre fault, VAH 5 Valledupar high, VH 5 Verdesia High. 2 km (1.2 mi)
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 443

TOC

10/21/15 11:26 AM
444 Sanchez and Mann

constrain the boundary conditions (heat flow, paleo- zone that curves from north-south in the south to east-
water depth, and SWIT) as well as the maximum burial west in the north (Bernal-Olaya et al., 2015a).
of the sedimentary section and source rock properties A significant positive gravity anomaly coincides
(TOC and HI). The section contains seven layers where with the SSMM and contrasts with the lower gravity TOC
two main unconformities show strata truncation by anomalies observed in adjacent basins of the LMV and
erosion. Estimation of erosion based on thermochro- Guajira area (Figures 1B and 5A). The high topographic
nology studies (Hernandez and Jaramillo, 2009; Vil- relief and gravity positive anomaly of the SSMM has
lagomez et al., 2011) and seismic interpretation was been interpreted by previous workers as an isostatic
introduced to the model, which allowed simulating the unbalance caused by recent uplift along the bound-
exhumation and evolution of structures and depocent- ing strike-slip faults of the Maracaibo Block (Case and
ers through six stages. The TR of the different source MacDonald, 1973; Ceron-Abril, 2008) (­ Figure 1B).
rocks was modeled for each stage. Gravity and magnetic modeling (Figure 4A) shows
the following ranges of values in potential fields
data: (1) a general gravity anomaly of 150-km (93.2-
Structural and Kinematic Modeling mi) half-wavelength and a maximum lateral contrast
of 300 mgal, where the highest values correspond to
Three WNW–ESE structural cross sections were con- the topographically elevated SSMM and the lowest
structed across the strike of the northern, central, and values correspond to the Caribbean–South ­American
southern parts of the 220-km-long (137 mi) CRB (Fig- Margin; (2) a general magnetic anomaly showing
ure 9). All three cross sections integrate sub-­surface 90- to 100-km (55.9- to 62.1-mi) half-wavelength with a
mapping based on 2-D seismic reflection lines tied lateral contrast of 150 nT onshore and 70 nT offshore;
to wells, gravity and magnetic modeling, and geo- and (3) onshore values correspond to a thin continen-
logic surface mapping (Gomez Tapias et al., 2007). We tal crust (ca 16–25 km [9.9–15.5 mi]) in the northwest-
loaded all these digital data into MOVE v. 2011­(Mid- ern Maracaibo Block or a high-density crust associated
land Valley) software to identify geometries, basin with Grenvillian metamorphic rocks that underlie
configuration, and structural styles and to quantita- much of the SSMM (Figure 4A).
tively restore the three cross sections in serial stages The crustal structure of northwestern South Amer-
from the recent to middle Cretaceous. ica constrained by the gravity model ­( Figure 4A)
These MOVE-based structural restorations follow includes a relatively shallow-dipping (ca 10–15°)
the principles of bed length balancing (Dahlstrom, ­Caribbean slab with a thickness of 6–7 km (3.7–4.3 mi)
1969) and utilize algorithms including “flexural-slip (Ewing et al., 1960; Mauffret and Leroy, 1997). The Car-
unfold” to restore folds and “fault parallel flow” to ibbean slab dips in a southeastward direction beneath
remove fault displacements. After each stage of res- the South American plate (van der Hilst and Mann,
toration, decompaction of the upper sedimentary sec- 1994; B ­ ernal-Olaya et al., 2015b). Shorter wavelengths
tion was conducted using standard values of surface seen in the gravity anomaly and the general trend of
porosity and the porosity-depth coefficient (Allen and ­magnetic anomaly indicate a depth to basement of
Allen, 2013). We also integrated thermochronologic 10–12 km (6.2–7.4 mi) both in the CRB and in the west-
and provenance results from previous workers into ern Maracaibo Basin. To the west, the gravity model
our structural and kinematic analysis (Shagam et al., constrains a sedimentary package thickness of 18 km
1984; Bayona et al., 2007; Hernandez and Jaramillo, (11.2 mi) in the accretionary prism formed along the
2009; Cardona et al., 2011a, b; Villagomez et al., 2011; thrust front of the subduction zone (South Caribbean
Ayala et al., 2012). deformed belt) (Bernal-Olaya et al., 2015b) (­ Figure 4A).
This deformed sedimentary wedge thins (ca 10–12 km
[6.2–7.5 mi]) along the frontal part of the accretionary
Results prism.

Crustal Structure
Gravity anomalies of the CRB
The Caribbean–South American Margin exhibits a
lower, regional gravity anomaly than the thinner and The regional gravity map shows two low gravity
higher density oceanic crust and oceanic plateau crust anomalies (ca 50–70 mgal) associated with the Cesar
underlying the subducting Caribbean plate to the north and Rancheria Subbasins to the south and to the north,
of the study area (Figure 1B). This gravity signature respectively (Figure 5A). The two subbasins are sepa-
shows the strongly arcuate shape of the subduction rated by a positive gravity anomaly corresponding

13880_ch16_ptg01_431-470.indd 444 10/21/15 11:26 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 445

to the Valledupar and V ­ erdesia highs (Figure 5B and contrast: (1) the northwestern magnetic low anomaly
C). This pattern of gravity highs and lows can be pro- with 20- to 30-km (12.4- to 18.6-mi) half-wavelength
jected eastward into the adjacent PR, which exhibits a corresponds to the northwestern (Santa Marta) and
gravity high in its central part with flanking gravity intermediate (Sevilla) belts of the SSMM (Tschanz et al., TOC
lows to the south and to the north with a contrast of 1974); and (2) an eastern anomaly high (Figure 5E) is
20–50 mgal ­between three parts of the PR. associated with the southeastern belt or Sierra Nevada
Using enhancements of residuals and tilt derivative Province (Tschanz et al., 1974). This eastern high mag-
(Figure 5B and C), we can use gravity to better define netic anomaly shows a half-wavelength of 20 km
the bounding faults and highs of CRB: (1) the CB is (12.4 mi). The contact between the central and southeast-
divided into two, preserved depocenters separated by a ern belt of the SSMM corresponds to the limit between
southern prolongation of the Verdesia High; (2) the RB these magnetic domains (Figure 5E) and also displays a
appears as an elongated gravity-low limited by NE–SW sharp edge on the tilt derivative map ­(Figure 5F).
lineaments corresponding to the Cesar lineament to the The CB is divided into two magnetic domains: (1)
northwest (Tschanz et al., 1974); (3) the Cerrejon thrust northern CB with a high anomaly and half-wavelength
fault to the southeast that is well expressed in the surface of 20 km (12.4 mi); some NE–SW and NW–SE linear
geology (Figure 2) because its fault plane juxtaposes the features are associated with short basement faults
Quaternary deposits of the RB with Jurassic and Creta- (­Figure 5F); (2) southern CB displays a lower anomaly
ceous strata of the PR (­ Kellogg, 1984; Montes et al., 2010; that continues to decrease southward to the SBF zone
Bayona et al., 2011); a gravity-low to the north of the PR is (Figure 5E). Between these domains there is no sharp
separated from the RB by a narrow high associated with and well-defined boundary (Figure 5F).
the Cerrejon thrust (Figure 5B and C); (4) the central CRB The RB is characterized by a high magnetic a­ nomaly
exhibits gravity highs (Valledupar and Verdesia) that are with a short half-wavelength of 10 km (6.2 mi) (Fig-
controlled by intersections of NE-SW and N-S lineaments ure 5E), which may be related to a shallower base-
related to basement faults; (5) gravity data reveal several ment block (Figure 5F). The PR is distinguished by a
discrete segments of the offshore Oca right-lateral, strike- magnetic low with a ca 20–30 km (12.4–18.6 mi) half-­
slip fault to the west and two clear segments of the SBF wavelength. This domain has a sharp boundary with
that correspond to the Santa Marta fault to the north the western RB while the boundary with the south-
and Bucaramanga fault to the south (Figure 5B and C). western CB is a more diffuse boundary (Figure 5F).
A well-defined gravity-low between these two seg-
ments may suggest a deeper depocenter associ-
ated with a normal faulting component; and (6) a Gravity and magnetic transects of the CRB
NWW-SEE local gravity-low in the SW flank of the
SSMM (Figure 5B) may be related to greater thickness Modeling of gravity and magnetic anomalies by con-
of the Jurassic batholith (Figure 1B) in comparison with structing six sections (located on the maps of Figure 5)
most other areas of the eastern SSMM where the meta- across the CRB and PR constrains the crustal structure
morphic belt may be thicker. of the region (Figure 6). The southern CB (Figure 6A
and B) has a crustal thickness of 25–30 km (15.5–18.6
mi) with a depth to the top of crystalline basement of
Magnetic domains of the CRB 7–10 km (4.3–6.2 mi). The gravity signature of the SBF
indicates a deeper depocenter to the southwest of the
Total magnetic field intensity illustrates two different SBF in the Lower Magdalena Basin (Figure 6A). A base-
domains in the CRB (Figure 5D). On the SSMM in the ment high in the western CRB near the PR is observed
west, there is an elongate magnetic anomaly low with with a thin lower and pre-Mesozoic sedimentary/meta-
a contrast of 300 nT in relation to a magnetic anomaly sedimentary section (Figure 6A, B, E, and F). In the
high over the CRB. East of the CRB on the PR a decreas- center of the southern CRB, the relatively flat-lying top
ing anomaly of 40–50 nT is noticeable. Reduction to the of basement (top of the upper crust shown in Figure 6)
pole of the magnetic data allows us to better locate the lacks significant highs or lows indicating uniform thick-
source of the magnetic anomalies (Figure 5E). Associat- ness in pre-Cretaceous (blue polygons in Figure 6) and
ing the magnetic anomalies with surficial geologic highs ­Cretaceous (green polygons in Figure 6) rocks.
and applying a tilt derivative helps to better define the The northern RB shows an eastward-dipping top
edges of underlying basement blocks (Figure 5F). basement surface (Figure 6C and D) with a depth range
According to these enhancements, we have iden- from 0.2 km (0.1 mi) to 5–7 km (3.1–4.3 mi). The modeled
tified magnetic domains based on the character of crust shows a general thickness of 20–25 km (12.4–15.5
the magnetic anomaly and qualitative wavelength mi). Some modeled sections (Figure 6C and D) indicate

13880_ch16_ptg01_431-470.indd 445 10/21/15 11:26 AM


446 Sanchez and Mann

the presence of rocks with metamorphic properties RB shows smaller reverse faults striking at high angles
toward the SSMM, which may be the downdip exten- to the Cerrejon fault. These reverse faults are truncated
sion into the subsurface of its easternmost metamorphic by the Eocene or the Miocene–Pliocene unconform-
belt. The RB shows a greater sedimentary thickness to ity (Figure 7A). Faulting is younger in the east where TOC
the north (Figure 6E) where the depth to basement is reverse faults cut through the Miocene–Pliocene
10 km (6.2 mi). The Oca right-lateral, strike-slip fault is unconformity and Quaternary deposits along
well defined by a deepening of the basement (Figure the western foothills of the PR (Figure 7A). In contrast,
6E) to 10–15 km (6.2–6.3 mi) with thickening of the sedi- Neogene deposits onlap the Eocene unconformity in
mentary and meta-sedimentary section. In the northern the western RB suggesting older deformation was in
PR, a thick section of pre-Cretaceous sedimentary and this area (Figure 7B).
metamorphic rocks is interpreted (Figure 6C, D, and F) Major pre-Eocene uplift took place in the western
as an ancestral Mesozoic rift basin that was inverted to CRB where Neogene deposits unconformably over-
generate the current topography relief of the PR. The lie the mid-Cretaceous section (Figure 7B), whereas
Valledupar High that separates the Cesar and Rancheria Late Cretaceous and Paleogene sections present in
Subbasins shows a modeled depth to basement of 1–2 the eastern RB are truncated by the Eocene uncon-
km (0.6–1.2 mi) (Figure 6E). formity (­ Figure 7A). Minor post-Paleogene uplift
occurred in the western RB where the Neogene sec-
tion is undeformed (Figure 7B). In contrast, erosional
Structure of the CRB truncations of Neogene strata against the Miocene–
Pliocene unconformity in the eastern RB (Figure 7A)
We summarize our seismic interpretation and struc- indicate that uplift was accompanied by faulting and
tural sub-surface mapping for five horizons of erosion. Major thickness variations for different units
the CRB that include: (A) near base of Cretaceous, (ca 0.8–1.6 km [0.5–0.9 mi]) from the Late Cretaceous
(B) intra-Cretaceous, (C) near top of Cretaceous, (D) to Paleogene (Figure 8G, H, and I) shows an east-
Eocene unconformity, and (E) Miocene–Pliocene uncon- ward migration of deformation toward the foothills
formity (Figure 8). Based on this information, we describe of the PR where large areas of erosion are present and
the major structures within the CRB and their evolution. recorded by the Eocene unconformity. A preserved
Subsurface and surface faults of the RB show two section of Neogene sedimentary rocks 0.2–0.8 km (0.1–
trends: (1) one fault trend parallels the PR (ca N30° E) 0.5 mi) thick is found in a narrow belt along the north-
and consists of eastward-dipping thrust and reverse ern RB (Figures 7A and 8I).
faults with a possible right-stepping relay arrange-
ment among the different fault segments; and (2) a
­second fault trend is parallel to N808 E linemanets Cesar Subbasin
and faults of the SSMM, which control two significant
highs (Valledupar and Verdesia Highs) in the center The relatively wider CB shows more structural com-
of the CRB (Figure 8A and B). Major structural lows in plexity consisting of basement-involved east-dipping
the CRB correspond to the La Loma syncline in the CB thrusts and some isolated westward-­dipping back-
and part of the footwall of the Cerrejon fault in the thrusts than the more narrow RB (­ Figure 7D and E).
RB ­(Figure 8A and B). Smaller structural highs in the The southern CB contains a prominent basement high
CRB are associated with local reverse faults. along its southern edge that may be associated with
recent left-lateral strike-slip deformation along the SBF
zone (Figure 7C). This high is mantled by Neogene
Rancheria Subbasin and Quaternary deposits, suggesting that the dip-slip
displacement of this fault segment decreased or ended
The structure of the northern RB consists of an eastward- by the early Miocene (Mora and Garcia, 2006).
dipping monocline controlled by the eastward-dipping The La Loma syncline forms a structural low with 5
Cerrejon thrust fault (Figures 2 and 7A), which presents km (3.1 mi) of structural relief and a complex fold axis.
a reverse throw of 2–3 km (1.2–1.9 mi) measured from We propose that there are two interfering fold axes: a
the base of the Cretaceous (Figure 8A). In addition, the main fold axis trending northeast and a secondary fold

Figure 7. Five, interpreted 2-D seismic lines in time (TWT) in the CRB showing main faults and sequences. Arrows illustrate
stratal terminations as either onlaps (purple) or truncations (blue). Ages are based on wells that are projected into the
­nearest seismic lines. Circled numbers next to wells are keyed to the burial history plots shown on Figure 10. Geologic
inset map indicates the positions of the five seismic lines. BA 5 Becerril anticline, LS 5 Loma syncline, SBF 5 Santa Marta–­
Bucaramanga fault, VAH 5 Valledupar High, VH 5 Verdesia High, VS 5 Los Venados syncline. 2 km (1.2 mi)

13880_ch16_ptg01_431-470.indd 446 10/21/15 11:26 AM


RB
A NW SE
Quaternary TWT
0.5
Neogene
Paleogene 1.0
A
Upper Cretaceous

13880_ch16_ptg01_431-470.indd 447
1.5
B

Figure 7. Continued
Lower Cretaceous
Pre-Cretaceous 2.0

Onlap strata termination


2.5
D Truncation termination
E Fault
3.0

2 km

C
CB RB
B SW NE
TWT
0.5

VAH
1.0

1.5

2.0

2.5

3.0
2 km.

9 CB
C Projected
TWT SW NE
0.5

1.0

1.5

2.0

LS
2.5

SBF
3.0

3.5

4.0

2 km
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 447

TOC

10/21/15 11:26 AM
448

D CB

13880_ch16_ptg01_431-470.indd 448
TWT NW 7 SE
0.5
Projected Quaternary
Neogene
1.0
Sanchez and Mann

VH BA Paleogene
1.5 Upper Cretaceous

Figure 7. Continued
Lower Cretaceous
2.0
Pre-Cretaceous
LS
2.5 Onlap strata termination
Truncation termination
3.0
Fault
3.5

2 km.

E CB
TWT NW 5 8 9 SE
0.5 Projected Projected

1.0

1.5
VS VH

2.0

2.5

3.0
LS

3.5

4.0

4.5

5.0 2 km
TOC

10/21/15 11:27 AM
A
TA
B C D E
CF

13880_ch16_ptg01_431-470.indd 449
VS
VH
BA

LS
meters

PERIOD EPOCH Ma Cesar


F G H I
Pliocene
10
Miocene Cuesta

Neogene
20

Oligocene 30

40
I
Eocene
50 Mirador (La Loma)

Paleogene
Paleocene 60 Cuervos
Barco
Barco
Ma 70 El Molino
H
Ca Socuy
80 Tres esquinas

Late
Sa
Co
Tu 90 G
La Luna
Ce
100 Aguas blancas
Al
110

Cretaceous
Lagunitas
Ap 120

Early
Ba 130 Rio Negro
F
He
Va 140
Be
150 La Quinta
Late
160

Middle
Macoita
170

Jurassic
180 Tinacoa
Early

25 km

meters
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 449

TOC

10/21/15 11:27 AM
450 Sanchez and Mann

Figure 8. Structural contour maps in depth (meters) for the Cesar–Rancheria Basin (CRB) for the following five interpreted
horizons that were mapped from a grid of 2-D seismic data tied to wells: (A) base of Cretaceous, (B) intra-Cretaceous,
(C) top of Cretaceous, (D) Eocene unconformity, and (E) Miocene–Pliocene unconformity. Thickness maps in meters are
shown for: (F) intra-Cretaceous to base of Cretaceous, (G) top of Cretaceous to intra-Cretaceous, (H) Eocene unconformity TOC
to top of Cretaceous, and (I) Miocene–Pliocene unconformity to Eocene unconformity. The stratigraphic inset chart indicates
the age range for the horizons and their mapped thicknesses and is based on the same sources used for the stratigraphic
­compilation shown on Figure 3. BA 5 Becerril anticline, CF 5 Cerrejo fault, LS 5 Loma syncline, TA 5 Tabaco anticline,
VH 5 Verdesia High, VS 5 Los Venados syncline. 10 km (6.2 mi)

axis trending northwest in the southern part of this The distribution of preserved depocenters indicates
structure (Figure 8A and B). This fold-axis pattern is the accumulation of 2 km (1.2 mi) and 3 km (1.9 mi) of
the likely result of either a multi-phase deformation or clastic sedimentary rocks during Early and Late Creta-
transpressional deformation (Ramsay and Huber, 1983). ceous, respectively (Figure 8F and G) in the La Loma
The La Loma syncline controls the greater preserved syncline of the eastern CB. Absent or thinner sections
thickness of Cretaceous and Cenozoic units because of (gray area in thickness maps) occur in the western and
significant pre-Eocene uplift to the north of CB where southern CB because of erosion. The preserved Pale-
Quaternary deposits unconformably overlie Early ocene is restricted only to the central part of La Loma
­Cretaceous and Jurassic rocks cropping out in the core syncline (Figure 8H), with a thickness of 0.6–1.2 km
of the Verdesia High (Figures 7C, 8G, and 8H). The Los (0.4–0.7 mi). This structure underwent little deforma-
Venados syncline to the west of the Verdesia High is tion as it remained in a footwall position during its
a smaller structural low with less structural relief than thrust-related uplift.
the La Loma syncline. The Los Venados syncline shows The Neogene section (Figure 8I) preserved to
a structural relief of 2.5 km (1.5 mi) (Figure 8A and B) the southwest of this subbasin shows a thickness of
and a complex fold axis. The eastern limit of La Loma ca 0.6–1.0 km (0.4–0.6 mi). Onlap strata relations toward
syncline is an eastward-dipping, NW-SE striking thrust the south (Figure 7C) indicate that pre-­Cretaceous rocks
with ca 1–1.5 km (0.6–0.9 mi) throw (­ Figure 8A and in the southern basement high may have been exposed
B). The hanging wall of this thrust includes the area and uplifted by the early Neogene. To the north, strata
between the PR front and La Loma syncline and shows underlying and dipping parallel to the Eocene uncon-
a complex fault pattern with a NE–SW main fault trend formity suggests tectonic quiescence during the early
and a secondary fault trend composed of E–W and Miocene followed by a late Miocene–Pliocene deforma-
NW-SE-striking fault populations (Figure 8A and B). tion pulse that affected the growth of the Verdesia High.
Most of the Late Cretaceous and upper sections are
truncated by the Eocene unconformity as seen on
­seismic lines (Figure 7C, D, and E). This unconform- Perija range and Sierra de
ity documents the age of ­Pre-Eocene shortening defor- Santa Marta massif
mation (Figure 8C and D). The section was removed
at this unconformity because the erosion during the Based on surficial geologic mapping (Gomez Tapias
Eocene uplift event was greater to the west (Figure 7D et al., 2007) (Figure 2), the PR is a fold-thrust belt
and E) and supportive of an older deformation linked bounded and internally deformed by reverse faults.
to the initiation of the SSMM uplift (Bayona et al., 2007, Along its western flank adjacent to the CRB, straight
2011; Cardona et al., 2011a, b). fault traces are interpreted as 35–40°, southeastward-
A gentle syncline south of the CB plunges to the dipping thrusts rooted at depth into basement and
southwest with fold axis trending NE–SW and is showing steeper dips of 50–60° dip (Figures 4B and
observed on the map of the Eocene unconformity 9C). The linearity of these reverse faults on surficial
­(Figure 8D), which documents post-Eocene deforma- geologic maps indicates a significant strike-slip com-
tion in this part of the basin. This syncline may have ponent on these faults.
been r­ e-deformed and displaced during the Neogene The Cerrejon fault system along the northwestern
by the left-lateral SBF (Villamil, 1999; Gomez et al., flank of the PR shows a relatively smaller dip angle of
2005a). For the Miocene and younger section, defor- 15–20° (Kellogg, 1984; Montes et al., 2010) than reverse
mation is concentrated to the east of the CRB where faults within the PR and controls an anticline cored by
faults at the foothills of the PR crosscut Quaternary Jurassic rocks (Figure 2). This relation suggests a ramp
deposits. The western CRB shows limited deformation of 50–60° dip that possibly involves basement (Figure
on a few, isolated faults (Figure 8D and E). 9A). Outcrops of sedimentary and meta-sedimentary

13880_ch16_ptg01_431-470.indd 450 10/21/15 11:27 AM


Section A
NW SSMM RB PR SE
0 km 10 20 30 40 50 60 A
MO-1 MCH-1

13880_ch16_ptg01_431-470.indd 451
0
km
CF ?
B
5

C
10

Section B
NW SE
SSMM CR PR
0 10 20 30 40 50 60 70 80

0
km
VH

5
?
?

10

Section C
NW CR SE
SSMM PR
0 10 20 30 40 50 60 70 80 90 100 110

C-H1X CO-1 PA-3


0
km

VS VH
5
? LS
?
?
10
SBF

Quaternary Upper Cretaceous Paleozoic


Miocene Lower Cretaceous Basement
Lower Eocene-Paleocene Pre-Cretaceous
Figure 9. Three structural cross sections of the Cesar–Rancheria Basin (CRB) are shown in depth (km) and integrate results from all 2-D seismic and well inter-
pretations from this study and published surface geologic maps compiled from previous works. Cross sections are constrained by dip data from surface geologic
maps and from dipmeter logs from wells—both measured sets of dips are shown by short, pink lines. Inset geologic map indicates the locations of the three,
cross sections. CF 5 Cerrejo fault, LS 5 Loma syncline, SBF 5 Santa Marta–Bucaramanga fault, VH 5 Verdesia High, VS 5 Los Venados syncline. 5 km (3.1 mi)
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 451

TOC

10/21/15 11:27 AM
452 Sanchez and Mann

Paleozoic rocks (Forero, 1970; Ujueta and Llinas, 1990) along the length of the CRB. In addition, the hydrocar-
associated with reverse faults within the PR may indi- bon TR is estimated using well information (Garcia et
cate a thick-skin structural style which controlled the al., 2008; Ayala, 2009) and displayed as a color overlay
exhumation of these older units. on the burial history graphs. TOC
The reactivation of pre-existing basement highs In general, the CRB shows an initial mild subsidence
formed during Mesozoic rifting within the PR are during the Cretaceous passive margin period consistent
indicated by the surficial linearity of fault trends, the with widespread thermally controlled s­ ubsidence and
large throw across basement-involved reverse faults, deposition of tabular, sedimentary units. ­During the
and abrupt and localized changes in the stratigraphic Cenozoic, there are two abrupt pulses of subsidence:
thickness of Jurassic, rift-related redbeds of the La one occurring during the ­Paleocene–early Eocene and
Quinta Formation (Miller, 1962) (Figures 1, 3). A prom- the other occurring during the early Miocene (Figure
inent, rectilinear topographic lineament parallels the 10). Both subsidence pulses are followed by two periods
NW–SE-striking El Tigre-Perija fault that in turn con- of exhumation: the first during the late Eocene–Oligo-
trols the Palmar and Totumo–Inciarte highs (Miller, cene and the second during the late Miocene–Pliocene.
1962; Shagam et al., 1984) in the eastern PR (Figure
2). The eastern flank of the PR shows Cenozoic-cored
synclines related to triangular zones (Figure 4B) that Rancheria Subbasin
may accommodate deformation beneath an upper
detachment and backthrust (Duerto et al., 2006). Northern areas of the RB (points 1 and 2 in Figure 2)
The SSMM is bounded by the left-lateral SBF to show rapid ­subsidence in the Paleocene (Figure 10A
the south and the right-lateral Oca fault (OF) to the and B) that result in the deposition of the thick, coal-
north. Both major faults are associated with conspic- bearing Cerrejon ­Formation, which includes high bitu-
uous NNE–SSW splay faults in the more intensively minous coals with vitrinite reflectance (Ro) values of
deformed apex area of the MB (Figure 2). 0.45% to 0.65% (Arango and Blandon, 2006). These
values indicate a maximum paleo-depth of 2.0–2.5 km
(1.2–1.5 mi) for this ­Paleocene section. After this sub-
Structural summary of the CRB sidence event, a strong exhumation pulse erodes some
of the Paleogene section. The absence of Neogene
The basin-wide structure of the CRB shows a shallower deposits may imply continuous uplift until the Plio-
basement in the northern RB (Figure 9A) at a depth of cene or deposition during the Neogene with an even
2–3 km (1.2–1.9 mi) and a deeper basement in the south younger uplift and erosional event (Figure 10B).
CB at a depth of 6–7 km (3.7–4.3 mi) (Figure 9B and C). Toward the southern RB (point 3 in Figure 1C), vit-
Shallower basement is also present along the western rinite reflectance shows values of 0.5% for the Late
CRB near the SSMM. The thickness of the pre-­Cretaceous Cretaceous Molino Formation and 0.65% for the Early
section increases southward (Figure 9C) and eastward in Cretaceous Lagunitas Formation, indicating maximum
the central part of the basin ­(Figure 9B and C). burial of 2 km (1.2 mi) for Early Cretaceous rocks (Fig-
The Cesar lineament, likely a basement fault observed ure 10C). In comparison to northern wells, the Neogene
on enhanced gravity maps (5B and C) separates the RB section is preserved and shows the lack of a late Mio-
from the SSMM but is not observed on 2-D seismic reflec- cene–Pliocene exhumation event. Using a pseudo-well
tion lines in this area (Figure 8A to E). Cenozoic defor- in the southern RB (point 4 in Figure 2) near the border
mation shows a pattern of eastward migration based with the CB, we propose that the absence of Late Creta-
on: (1) the undeformed Neogene section in the western ceous to Paleogene strata may record to the uplift of the
CRB; (2) faults cutting younger P ­ liocene-Pleistocene Valledupar High, after which continuous subsidence
rocks in the eastern CRB while rocks of this same age are and deposition of Neogene and Quaternary section
undeformed in the western CRB; and (3) the presence of records a period of tectonic quiescence (Figure 10D).
a major erosional event in the western CRB during the
Paleogene and the presence of a major erosional event
in the eastern CRB during the Neogene (Figure 8F to E). Cesar Subbasin

The southwestern CB (point 5 in Figure 2) shows an


Basin Analysis absence of the Paleogene section that is interpreted
as erosion ­during an Eocene–Oligocene exhumation
We constructed burial history graphs based on seven pulse followed by no significant deformation or dep-
wells, one pseudo-well, and a published stratigraphic osition ­during the Neogene and Quaternary (Figure
column based on outcrop data (Figures 2 and 10) to 10E). Thermochronological analysis for the eastern CB
summarize the major subsidence and uplift events indicates an ­accelerated exhumation onset by 10–15

13880_ch16_ptg01_431-470.indd 452 10/21/15 11:27 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 453

Ma for Early Cretaceous and Late Jurassic units using indicate a maximum burial of 5.0–5.5 km (3.1–3.4 mi)
AFT (Hernandez and ­Jaramillo, 2009). for the lower Cretaceous. Such burial would lead to
We calibrated our burial history diagrams of this complete resetting of AFT ages. The Paleogene and
area (points 6 and 7 in Figure 2) with these available early Neogene sections may have undergone other TOC
thermochronological ages and with vitrinite reflec- exhumation events (Figure 10F, G and H), or may
tance that shows values of 1.3–1.5% for the Late Cre- have undergone a single, prolonged subsidence event.
taceous La Luna Formation, 0.6–0.9% for the Early Given the burial history of other areas in the CRB, we
Cretaceous Cogollo Group, and 2.1% for the Early favor the interpretation of a minor Eocene–Oligocene
Cretaceous Rio Negro Formation. All vitrinite values exhumation event ­(Figure 10C and E).

A Jurassic Cretaceous Paleogene Neogene


B Jurassic Cretaceous Paleogene Neogene
Conj. Calc. Conj. Calc.
Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio

Cerrejon
Cerrejon Manantial
Hato Nuevo
Manantial Molino
Hato Nuevo
Molino La Luna
La Luna Cogollo
Cogollo La Quinta
La Quinta

1 2
Phase 1
Phase 1
Phase 2 Phase 2

C Jurassic Cretaceous Paleogene Neogene


D
Jurassic Cretaceous Paleogene Neogene
QT.
Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio
Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio

Conj. Cong. QT./Plio.


Conj. Calc. Cuesta
Cerrejon La Luna
Manantial Cogollo
Hato Nuevo Rio Negro
Molino La Quinta
La Luna
Cogollo
La Quinta

Phase 1

3 Phase 2
4 Phase 1
Phase 2

E Jurassic Cretaceous Paleogene Neogene


FJurassic Cretaceous Paleogene Neogene
QT.
Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio

Cuesta QT.
Molino Molino
La Luna
Cogollo La Luna
Rio Negro
La Quinta
Cogollo

Rio Negro
La Quinta

5 6
Phase 1
Phase 1
Phase 2 Phase 2

Figure 10. Burial history graphs for nine different locations within the CRB based on either well data located on Figure 2 or
previously published stratigraphic data as summarized on Figure 3. Main subsidence events at each location are shown by
blue arrows; uplift events are shown by red arrows. Vitrinite reflectance and thermochronologic data were used to constrain
the amount and timing of subsidence. Calculated hydrocarbon transformation ratio for source rocks is displayed as a colored
overlay on the burial history graphs. See text for discussion.

13880_ch16_ptg01_431-470.indd 453 10/21/15 11:27 AM


454 Sanchez and Mann

G Jurassic
Upper Jurassic Lower Cretaceous
Cretaceous
Upper Cretaceous Paleocene
Paleogene
Eocene Oligocene
Neogene
Miocene Plio
QT.
I Jurassic Cretaceous Paleogene Neogene
Molino Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio
QT.
La Luna
Molino
Cogollo La Luna
TOC
Cogollo
Rio Negro
La Quinta Rio Negro
La Quinta

7 Phase 1
8 Phase 1
Phase 2
Phase 2

H Cretaceous Paleogene Neogene


Jurassic
Upper Jurassic Lower Cretaceous Upper Cretaceous Paleocene Eocene Oligocene Miocene Plio

QT.
Cuesta
Jagua
Cuervo/
Barco
Major uplift event

Molino Major subsidence event


La Luna

Cogollo
Hydrocarbon transformation
ratio
Rio Negro
La Quinta

9 Phase 1
Phase 2

Figure 10. Continued

The southern continuation of the Verdesia High in Group). The ratio shows values of >50% t­ oward the
the CB (point 8 in Figure 2) shows a maximum paleo- northern RB and eastern CB where source rocks were
depth of 2.5–3.0 km (1.5–1.9 mi) based on vitrinite buried to depths of 3–4 km (1.9–2.5 mi) (Figure 10A, B,
reflectance (Ro) of 0.6–0.7% for the Late Cretaceous and G). In the center of the La Loma syncline, source
Molino Formation, 0.7–0.8% for the Late C ­ retaceous rocks also reach high TR at burial depths of 3–4 km
La Luna Formation, and 0.8–0.9% for the Early Cre- (1.9–2.5 mi) (Figure 10H). In other areas, hydrocarbon
taceous Cogollo Group. The absence of the Paleogene TR is smaller (<50%) ­because of shallower burial of
and Neogene sections requires high exhumation and source rocks as seen in the southern RB at 2 km (1.2
deformation during the Eocene–­Oligocene and late mi) deep (Figure 10C), western CB at 2.5 km (1.5 mi)
Miocene–Pliocene convergent events (Figure 10I) that deep (­Figure 10E), or close to the Valledupar or Verd-
accompanied the uplift of the Verdesia High. The La esia Highs at a depth of 2–3 km (1.2–1.9 mi) (Figures
Loma syncline (point 9 in F ­ igure 2) shows a well- 10D, F, and I).
preserved Paleogene and Neogene section, indicat- According to the observed burial history, the lack of
ing reduced exhumation (Figure 10H) because of its Late Cretaceous and Paleogene sections suggests exhu-
distant location from overthrusting mountain fronts mation during late Eocene–Oligocene or Miocene–­
bounding the CRB. Vitrinite reflectance of 0.4–0.5% for Pliocene deformation events – rather than depocenter
Paleocene Cuervos Formation, 0.5–0.6% for the Late distribution exclusively controlled by pre-existing
Cretaceous Molino Formation, and 0.7–0.8% for the structural highs. As in the thickness maps (Figure 8F
La Luna Formation allows estimates of a maximum to I), the burial history of the Neogene section reveals
­burial of 3.5–4.0 km (2.2–2.5 mi) for Early Cretaceous. a change in the distribution of depocenters linked to
patterns of west-to-east exhumation and related to the
migration of Cretaceous–Paleogene depocenters.
Maturation and distribution of source pods A 2-D basin model across the southern CB based on
a structural cross section (Figure 9C) was calibrated
Hydrocarbon TR records the maximum burial of with three burial history diagrams and summarizes
source rocks (organic-rich levels in the Cretaceous the development of Paleogene and Miocene depo-
Molino and La Luna Formations and the C ­ ogollo centers (Figure 11A to F). An initial uplift of the area

13880_ch16_ptg01_431-470.indd 454 10/21/15 11:27 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 455

NW SSMM CB PR
SE
10 km 20 30 40 50 60 70 80 90 100 110
5 8 9
0

5 A TOC

10
km
0 Ma
0

5
B
10
km 5 Ma
0

C
5

10
km
23 Ma
0

10 D
15
40 Ma
0

10 E
15
65 Ma
0

10 F
15
90 Ma
Transformation
Stratigraphy and Source Rocks Ratio (%)
Introduced in the Model
Quaternary
Early Cretaceous
Miocene

Paleocene– Pre-Cretaceous
Middle Eocene
Basement
Source rock
Late Cretaceous

Figure 11. Two-dimensional basin model in depth across the southern Cesar Subbasin (CB), the easternmost part of the
­ ierra de Santa Marta massif (SSMM), and the western part of the Perija range (PR) based on structural cross section C
S
shown on Figure 9. The model shows six different stages (A) to (F) with the oldest stage A being in the late Cretaceous
(ca 90 Ma). Smaller, inset sections for each stage summarize the calculated, hydrocarbon transformation ratio for that time
period. Three burial history graphs shown in Figure 10 were used to calibrate this 2-D basin model. 5 km (3.1 mi)

13880_ch16_ptg01_431-470.indd 455 10/21/15 11:27 AM


456 Sanchez and Mann

adjacent to the SSMM by the Paleocene (Bayona et al., tilting event that was controlled by the exhumation
2007; Villagomez et al., 2011) was followed by a sig- of the SSMM to the west (Montes et al., 2005b). The
nificant exhumation event by late Eocene–­Oligocene growth of the Verdesia High began by the Eocene–Oli-
that induced erosion to the west and a younger, late gocene deformation event, which is associated with TOC
Miocene–Pliocene event that induced major ero- reactivation of previous, basement faults (Figure 12B3
sion to the east. Hydrocarbon transformation ratios and B4). After Miocene time, deformation was mainly
(insets in Figure 11A to F) indicate that higher values focused in the eastern CRB where major structures
of transformation (ca 80%) are reached by deep, mid- developed in the PR and eastern CB and produced
dle Eocene burial of Cretaceous source rocks in the La erosion of the Paleogene and most of the Late Cre-
Loma syncline. taceous sections. ­C onsequently, major depocenters
migrated to the western CB (Figure 12B5 and B6). The
shortening of the CB section is around 8%.
Kinematic Evolution of the CRB The 110-km-long (68.3-mi) southern section (­ Figure
9C) across the CB presents a total shortening of 6%
Two-dimensional structural restorations for the three (Figure 12C). The Paleogene section is thickest (ca 2
constructed sections across the strike of the CRB show km [1.2 mi]) in the area of the La Loma syncline and
the evolution of thrust deformation and its ­control Verdesia High during the middle Eocene. In the west-
on CRB sedimentation (Figure 9). For the northern ern CRB, the Paleogene section is thinner or absent
cross section (Figure 9A), the kinematic evolution because of the uplift of the SSMM (Figure 12C2). The
based on the basin model and published thermochro- Late Eocene–early Miocene deformation event would
nological data (Shagam et al., 1984; ­Villagomez et al., produce higher uplift to the west that would control
2011) indicates that the 60-km-long (37.3-mi) ­section the exposure of the Late Cretaceous units and remnant
underwent a total shortening of 13% (Figure 12A). Paleogene rocks to the east (Figure 12C3 to C5). During
Key observations seen on the restorations of the the Miocene, convergent deformation was accommo-
northern section shown in Figure 12A include: (1) the dated by thrusts to the east of the La Loma syncline,
Cretaceous is a period of passive margin subsidence and deposition took place in the western CRB above
with little or no deformation of tabular, sedimen- the angular unconformity produced by the Eocene–
tary units; underlying the passive margin section are Oligocene shortening event (Figure 12C6). After the
wedge-shaped, Jurassic sedimentary units controlled Miocene, continuous uplift of the PR exposed Early
by both normal fault and the locations of pre-exist- Cretaceous and Jurassic rocks at the surface (Figure
ing Paleozoic basement highs (Figure 12A1); (2) the 12C7).
­Paleogene section of the CRB was deposited and then
erosionally removed in the west by the onset of uplift
along the SSMM during the Paleocene–Eocene period Summary of structural restorations
(Figure 12A4); (3) a strong deformation event occurred
between the late Eocene–lower Miocene and resulted Our restorations suggest that shortening of the
in the erosion of the Paleogene section and part of northern CB is greater than the central and south-
the Cretaceous section to the west (Figure 12A4); (4) ern CRB (Figure 12B) and is responsible for the
continued deformation of the PR during the Miocene exhumation of Jurassic rocks in the Verdesia High
restricted the depocenter of the CRB to the axis of the in comparison to the southern CB where this same
RB (Figure 12A5); (5) a strong, Late Miocene–Plio- high is buried by recent deposits (Figure 12C).
cene deformation pulse led the uplift of the PR, which Minor shortening in the RB reflects its simpler
eroded the entire Paleogene and Cretaceous sections ­s tructure compared to more complex thrusting of
and part of the Jurassic section (Figure 12A6); (6) this the CB. Along the western PR front, shortening
last stage of deformation accommodated most of the increases to the north, which may be a consequence
total shortening in the CRB: most of this shortening of the El Tigre–Perija strike-slip fault whose NE–SW
was absorbed in the PR as the total shortening meas- trace is closer to CB than the RB (­ Figure 2) and may
ured in the RB is only 6%. have overprinted the southern PR with a strike-slip
The section to the north of the CB crosses the Verd- component. The general low shortening seen on all
esia High, cored by outcrops of the Jurassic La Quinta three cross sections is a consequence of the dominant
Formation (Figure 9B). This 90-km-long (55.9 mi) cross thick-skin structural style of the CRB with high-angle
section shows a total shortening of 11% (Figure 12B), faults with dips of 35–60° that are rooted into base-
with an initial uplift during the Paleocene–Eocene ment and show few significant, horizontal detach-
(Figure 12B2) that accompanied an eastward, regional ment levels.

13880_ch16_ptg01_431-470.indd 456 10/21/15 11:27 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 457

A. Kinematic restoration of Section A


NW SE
SSMM RB PR
0 10 20 30 40 50 60

MO-1 MCH-1
TOC
0
CF
A

5
B
Recent
10 Shortening: 8.9 km (12.7%) C
Only for Rancheria Sub-basin: Shortening: 2.4 km (5.6%) 1

Pliocene–Miocene (~5 Ma)


10 Shortening: 2.9 km (4.2%)
Only for Rancheria Sub-basin: Shortening: 1.1km (2.5%) 2

Miocene–Oligocene (~23 Ma)


10 Shortening: 1.6 km (2.4%)
Only for Rancheria Sub-basin: Shortening: 0.9 km (2.2%) 3

Oligocene–Eocene
10 Shortening: 1.1 km (1.6%)
Only for Rancheria Sub-basin: Shortening: 0.9 km (2%) 4

Eocene (~45 Ma)


10 Shortening: 1 km (1.4%)
Only for Rancheria Sub-basin: Shortening: 0.06 km (0.1%) 5

10 Middle Cretaceous (~90 Ma)


Shortening: 0 km 6
Figure 12. Serial structural restoration for cross sections A, B, and C across the northern CRB shown in Figure 9. Shortening
values were calculated at several different times during the Late Cretaceous and Cenozoic. Published thermochronologic data
summarized on Figure 4 were used to calibrate the five different stages of the structural restoration. CF 5 C
­ errejo fault,
CR 5 Cesar Subbasin, LS 5 Loma syncline, PR 5 Perija range, RB 5 Rancheria Subbasin, SBF 5 Santa Marta–­Bucaramanga
fault, SSMM 5 Sierra de Santa Marta massif, VH 5 Verdesia High, VS 5 Los Venados syncline. Key to colored rock units is
shown on Figure 9.

13880_ch16_ptg01_431-470.indd 457 10/21/15 11:28 AM


458 Sanchez and Mann

B. Kinematic restoration of Section B


NW SE
SSMM CR PR
0 10 20 30 40 50 60 70 80

TOC
0

VH

Recent
Shortening: 10.3 km (10.6%)
1
10
Only for Cesar Sub-basin: Shortening: 5.5 km (8%)

Miocene–Pliocene
10
Shortening: 5.3 km (5.4%)
Only for Cesar Sub-basin: Shortening: 3.4 km (4.9%) 2

Miocene–Oligocene (~23 Ma)


10
Shortening: 3.3 km (3.4%)
Only for Cesar Sub-basin: Shortening: 0.9 km (1.4%) 3

Oligocene–Eocene?
10 Shortening: 1.1 km (1.2%)
Only for Cesar Sub-basin: Shortening: 0.5 km (0.8%) 4

Eocene (~45 Ma)


10 Shortening: 0.3 km (0.3%)
Only for Cesar Sub-basin: Shortening: 0.01 km (0%) 5

10
Middle Cretaceous (~90 Ma)
Shortening: 0 km 6
Figure 12. Continued

13880_ch16_ptg01_431-470.indd 458 10/21/15 11:28 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 459

C. Kinematic restoration of Section C


NW SE
SSMM CB PR
0 10 20 30 40 50 60 70 80 90 100 110

0
C-H1X CO-1 PA-3
TOC
VS VH
5
SBF LS

10
Recent
Shortening: 6.5 km (5.5%) 1

10
Pliocene–Miocene (~5 Ma)
Shortening: 5.3 km (4.6%) 2

10
Miocene–Oligocene (~23 Ma)
Shortening: 4.3 km (3.7%) 3

10
Oligocene–Eocene?
Shortening: 3.3 km (2.8%) 4

10
Eocene (~37)
Shortening: 2.3 km (2%) 5

10
Eocene (~45 Ma)
Shortening: 1 km (0.9%) 6

10
Middle Cretaceous (~90 Ma)
Shortening: 0 km 7
Figure 12. Continued

13880_ch16_ptg01_431-470.indd 459 10/21/15 11:28 AM


460 Sanchez and Mann

Discussion late Miocene to recent effects of the Panama Arc colli-


sion as proposed by Vargas and Mann (2013).
Tectonic Events Affecting the CRB
TOC
Two major deformational phases affecting the CRB Correlations with the Maracaibo Basin
were identified and linked to 3–7 km (1.9–4.3 mi) of of Venezuela
regional tectonic shortening observed in the CRB: (1)
an Early-Middle Eocene convergent phase observed The Maracaibo Basin, which is physically separated
in the western CRB controlled the general eastward from the western CRB by the barrier of the PR ­(Figure 1B
tilting of the Mesozoic section; this convergent phase and 2), shows a more complete, less eroded, and less
has been related by us and most previous workers to deformed Cenozoic section (Parnaud et al., 1995; Aude-
the accretion of the Great Arc of the Caribbean (Car- mard and Serrano, 2001) (Figure 3). The Maracaibo
dona et al., 2011a; Villagomez et al., 2011); and (2) a Basin is characterized by wedge-shaped, sedimentary
late Miocene–Pliocene phase characterized by large-­ packages recording tectonic controls by arc-continental
displacement thrusting in the eastern CRB driven by collision between South American and the Great Arc
late Miocene collision of the Panama Arc (Mann and of the Caribbean including the formation of an asym-
Burke, 1984; Burke, 1988; Eva et al., 1989; Taboada, metrical, Eocene foreland basin (Lugo and Mann, 1995;
2000; Vargas and Mann, 2013). This Miocene defor- Mann et al., 2006; Escalona and Mann, 2011).
mation event accompanied the northward ”tec- Evidence for this same Eocene convergent and fore-
tonic escape” of the MB that remains active today land basin event in the CRB and PR region to the west
(Colmenares and Zoback, 2003) and includes the include: (1) the formation of the NW–SE-trending
period of Miocene to recent uplift of the Sierra de Macoa–Totumo arch in the eastern PR with truncation
Santa Martha massif (SSMM) (Figure 12). of the Paleocene Marcelina and Guasare Formations
and underlying Cretaceous rocks (Kellogg, 1984); (2)
along the eastern flank of the PR, the Eocene La Sierra
Ages of the two convergent events Formation, Paleocene, Cretaceous, and pre-Cretaceous
rocks are all truncated beneath the early Miocene
We propose that the Cenozoic deformation is driven by Fausto Group and record the deformation and rapid
the diachronous collision of the Great Arc of the Carib- uplift of this flank of the PR during the Oligocene and
bean that has migrated from west to east as observed Eocene (Kellogg, 1984)
by previous workers in the Middle Magdalena Valley Our interpretations in the CRB show a more struc-
Basin (MMV) and Eastern Cordillera (Gomez et al., turally elevated basement and much thinner sedimen-
2005b; Parra et al., 2009; Mora et al., 2010). The onset tary section in comparison to the thicker sedimentary
of deformation and uplift has been proposed to be section and more depressed crystalline basement of
Paleocene in the SSMM (Villagomez et al., 2011) and the western Maracaibo Basin (Figures 4B and 13B). The
Eastern Cordillera (Villamil, 1999; Cortes et al., 2006; structure of the CRB reveals a deeper level of exhuma-
Parra et al., 2012; Mora et al., 2013), Eocene in the PR tion and suggests the proximity of the CRB to the arc-
(Kellogg, 1984; Bayona et al., 2007), and Oligocene in continental collision zone by the Paleocene–Eocene, as
the northern Andes (Merida Andes) (Bermudez et al., well as the possible contribution of right-lateral shear-
2010). Within the CRB, we also observe a progressive ing during the accretion of the Great Arc of the Car-
deformation of younger Pliocene–Pleistocene rocks to ibbean (Burke, 1988; Kennan and Pindell, 2009). The
the east and westward migration of preserved Pale- strongly curved shape of the subduction margin of
ocene and Miocene depocenters (Figures 8, 11, and 12). northwestern South America may have induced rapid,
The presence of Paleocene limestone units deposited lateral changes as the result of radial normal ­faulting
during a period of tectonic stability along the eastern of the overriding forearc area (Feuillet et al., 2002;
PR (Kellogg, 1984) also supports our proposal for Bernal-Olaya et al., 2015c) (Figure 1B).
­C enozoic, west-to-east migration of deformation
­related to the Great Arc of the Caribbean collision.
The rapid, 10 Ma and younger exhumation event Structural style of the CRB
seen in SSMM (Cardona et al., 2011b; Villagomez et al.,
2011), eastern CRB, and PR (Kellogg, 1981, 1984; Shagam We propose a dominantly basement-involved struc-
et al., 1984; Hernandez and Jaramillo, 2009) accom- tural style for the CRB based on our interpretations of
panies rapid late Miocene-Pliocene exhumation and structural cross sections through the CRB that include
deformation in the Venezuelan Andes (Audemard and the following features: (1) linear gravity and magnetic
Audemard, 2002; Bermudez et al., 2011) and in the East- trends that suggest that major faults are thick-skinned
ern Cordillera (Parra et al., 2009). This synchronous and and penetrate into basement (Figure 5B, C, and F); (2)
regional deformation is supportive of the widespread, a high dip angle of 35–60° for the main strike-slip and

13880_ch16_ptg01_431-470.indd 460 10/21/15 11:28 AM


13880_ch16_ptg01_431-470.indd 461
A 1400
1200
1000
800
600

Age (Ma)
400
200
97000 107000 117000 127000 137000 147000 157000 167000 177000 187000
0

B SSMM RB PR MB
0 km 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190

MO-1 MCH-1
42 47
0
km AFT:
43.4±5.4 Ma

5 CF
AFT:
10 TF 2.6±0.4 Ma
ZFT:
117±19 Ma
15 NW SE
Geochronology
Plio/Pleistocene Paleocene Jurassic/Triassic U-Pb
Late Miocene
Late Cretaceous Paleozoic Rb / Sr
Early Miocene
Oligocene/Eocene K/ Ar , Ar / Ar
Early Cretaceous Basement
Depositional age

Figure 13. Regional cross section B shown as the red line on the inset map integrating stratigraphic and structural information for the Rancheria Subbasin (RB), Perija
Range (PR), and western Maracaibo Basin (MB). The regional section of the MB, Venezuela, was modified from a previous cross section by Mann et al. (2006). The
El Tigre (TF) and Cerrejon (CF) faults are main features in the PR. Thermochronologic data from Shagam et al. (1984) and Villagomez et al. (2011) are posted along
in the section. Geochronologic data (A) were compiled from Tschanz et al. (1974), MacDonald and Opdike (1984), Bayona et al. (2010), and Cardona et al. (2011a).
5 km (3.1 mi)
Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 461

TOC

10/21/15 11:28 AM
462 Sanchez and Mann

thrust faults based on their linearity on surface geo- Nova et al., 2012). The observed SSMM clockwise rota-
logic maps and in the subsurface (Figures 2 and 7); tion of 12°–18° (Bayona et al., 2010; Nova et al., 2012)
and (3) the involvement of pre-Cretaceous rocks in the may indicate differing deformation history for the
hanging wall of major thrust faults (Figures 4B and SSMM and PR. In our study area of the CRB, our cross TOC
13B). sections do not support any significant differences in
Low shortening values (<10%) along the CRB sup- the structural style or amount of pre-Cenozoic defor-
port our proposed, thick-skinned structural style and mation observed in the PR and SSMM (Figure 12).
are consistent with the calculations of 6-km (3.7-mi) The basement of the SSMM province has been pre-
shortening for the eastern flank of the PR (Duerto et viously dated as Grenville in age (1200–1000 Ma)
al., 2006) and 58–80 km (36–48.7 mi) for the entire East- (Restrepo-Pace et al., 1997; Cordani et al., 2005; ­Cardona
ern Cordillera (Mora et al., 2008; Teson et al., 2013) Molina et al., 2006), whereas the PR and Maracaibo
(Figure 1). Alternative thin-skinned models by previ- Basin are underlain by Paleozoic (500–310 Ma) base-
ous workers propose low-angle thrusts with displace- ment of igneous and metamorphic rocks (Feo-Codecido
ments of 16–26 km (9.9–16.1 mi) along the Cerrejon et al., 1984) (Figure 4D). Our regional gravity and mag-
fault and total shortening of 46–56 km (28.6–34.8 mi) netic modeling (Figure 4A) suggests a higher density
for the RB and PR (Kellogg and Bonini, 1982; Kellogg, and lower magnetic susceptibility crust for the SSMM in
1984; Montes et al., 2010) or 140–190 km (86.9–118 mi) comparison to the upper crust of the PR and Maracaibo
for the Eastern Cordillera (Dengo and Covey, 1993; Basin. This higher gravity and lower magnetic anomaly
Roeder and Chamberlain, 1995). Differences in short- (Figures 1B and 5) indicates that a major crustal bound-
ening along the CRB on our three cross sections show ary separates these two different types of basement.
a slightly higher shortening in the central part of the Paleozoic meta-sediments cropping out in the PR are
CRB (8%, Figure 12B) in comparison to the northern absent in the SSMM and generate uncertainty in the
and southern parts where shortening is 5–6% (Figure construction of our cross sections (Figure 9).
12A and C). Greater shortening of the central CRB is
consistent with the presence of the Valledupar base-
ment high in the central CRB (Figures 5B, C and 6E). Basin Evolution and Paleogeography
Our data does not support the model by Montes
et al. (2010) that invokes large rotationally-induced We summarize our understanding of the interac-
deformation of the SSMM with consequent high short- tions between tectonics and basin formation in the
ening in the adjacent RB. Studies in the southwestern CRB region of northwestern South America using a
Venezuelan Andes (Figure 1B) propose shortening of series of paleo-geographic maps for four time periods:
10 km (6.2 mi) (Duerto et al., 2006) for the northern Late Cretaceous, Paleocene, Oligocene, and middle
flank, 12 km (7.5 mi) increasing to 45 km (27.9 mi) to ­Miocene (Figure 14). These paleo-geographic maps
the northeast ­(Audemard, 1991), 20–30 km (12.4–18.6 were created using well and outcrop information pre-
mi) (De Toni and K ­ ellogg, 1993), 40–50 km (24.8–31 sented in this chapter and by compiling previously
mi) ­(Audemard and Audemard, 2002), or 50–60 km published paleo-geographic maps from other authors
(31–37.3 mi) (Colletta et al., 1997). Such amounts of (Villamil, 1999; Mann et al., 2006; Escalona and Mann,
observed deformation are significantly less than the 2011; Ayala et al., 2012).
shortening amount predicted by 50–75° clockwise
rotation of the SSMM Block (Montes et al., 2005a).
Paleomagnetic analyses indicate a vertical-axis, Major paleo-geographic stages
clockwise rotation for Jurassic and Cretaceous rocks
of 40–50° for the entire PR with initial magnetization Five main stages are observed in the paleo-geographic
during the pre-deformation period (Gose et al., 2003; evolution of northwestern South America: (1) slow,

Figure 14. Paleo-geographic maps showing the evolution of northwestern South American at four critical time intervals
­ odified from Villamil (1999), Mann et al. (2006), Escalona and Mann (2011), and Ayala et al. (2012). All maps show the
m
modern outline of the Cesar-Rancheria Basin (CRB) in blue. Time intervals include: (A) Late Cretaceous (ca 80 Ma). D­ uring
this time the area of northern Colombia and the CRB was a broad, north-facing passive margin. The CRB, Guajira, and
­Maracaibo Basins formed a ­continuous basin in this broad, passive margin setting. (B) Paleocene (ca 60 Ma). The Great
Arc of the Caribbean began to ­collide with the passive margin and a foreland basin developed along the collisional zone.
(C) Oligocene (ca 30 Ma). The Great Arc collision migrates eastward and much of the area became shortened and emer-
gent. (D) Pliocene (ca 5 Ma). The ­collision of the Panama Arc west of the study area activated the northward motion of the
­Maracaibo Block that included the CRB and accelerated the topographic elevation and widening of the northern Andes.
BF 5 Burro Negro fault, OF 5 Oca fault, LN 5 Lara Napes, PMV 5 proto-Magdalena Valley, PR 5 Perija range, SBF 5 Santa
Marta–­Bucaramanga fault, SM 5 Santander massif, SSMM 5 Sierra de Santa Marta massif, TF 5 El Tigre fault.

13880_ch16_ptg01_431-470.indd 462 10/21/15 11:28 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 463

A B
Late Cretaceous: ~80 Ma Paleocene: ~60 Ma
TOC

Guajira salient
Shelf edge Guajira
Maracaibo basin
Shelf edge

era
Merida arch

ill
Cord
Maracaibo basin

tral
Cen
C D
Oligocene: ~30 Ma Pliocene: ~5 Ma

Falcon basin
OF
OF
SSMM
SB

SSMM
F

PR
BNF Magdalena fan
SB

LN TF
F
Romeral suture

PR
illera
Western Cord

PMV SM Merida Andes


Central Cordillera

Middle-Outer shelf Caribbean plateau Alluvial fans

Upper shelf-deltaic Foreland basin Turbidites

Fluvial-Coastal plain Control points


Proto-Caribbean (wells)
Carbonate shelf Emergent area Study area
(CRB)
GAC (Great arc of the Caribbean)

Accretionary prism

13880_ch16_ptg01_431-470.indd 463 10/21/15 11:28 AM


464 Sanchez and Mann

thermally driven, Cretaceous subsidence in a passive Comparison of CRB deformational events


margin setting followed Jurassic rifting (Figure 14A); to other, neighboring basins
this passive margin phase is characterized by tabular
and mixed carbonate-clastic facies of the middle to In the Middle Magdalena Basin (MMB) south of the TOC
outer shelf (Pindell and Tabbutt, 1995; Villamil, 1999; CRB burial history plots show an important exhuma-
Cediel et al., 2003); (2) an abrupt and rapid Paleocene tion episode occurring by middle Eocene (Gomez et
subsidence, previously documented by Bayona et al. al., 2005b) that is synchronous with Eocene deforma-
(2011) and Ayala et al. (2012), is attributed to flexure of tional events we propose for the CRB (Figure 10). This
the continental margin of northwestern South Amer- Eocene event is not observed in the LMV Basin to the
ica in response to the late Cretaceous-Paleogene col- east of the CRB (Bernal-Olaya et al., 2015c).
lision of the Great Arc of the Caribbean (Cardona et Late Miocene uplift event in the CRB also correlates
al., 2011a); this collisional event generated significant with the onset of exhumation for the western foothills
thickness changes in the sedimentary section because of the Eastern Cordillera (Mora et al., 2010; Sanchez
of the fault-bounded isolation of the CRB that included et al., 2012) (Figure 1). A widespread uplift episode
a period of very rapid, Paleocene (ca 58-60 Ma), flu- along the LMV Basin also produced a very prominent,
vial to coastal-plain deposition of the coal-bearing Late Miocene angular unconformity (Bernal-Olaya et
Cerrejon Formation of the northern CRB (Ayala et al., al., 2015c). The similarity in basinal tectonic events
2012) ­(Figure 14B); (3) Late Eocene–Oligocene exhu- between the MMB and the CRB suggests a shared
mation of the CRB and neighboring Maracaibo Basin stratigraphic connection that has been offset by the
to the west is related to isostatic rebound of South 110-km (68.3 mi), left-lateral displacement of the SBF
American continental crust following its collision with system during the Neogene (Campbell, 1965). This off-
the Caribbean Arc (Escalona and Mann, 2011); dur- set is driven by the “tectonic escape” of the Maracaibo
ing this deformation event, no deposition took place Block that was induced by the late Miocene collision
with the formation of large erosional unconformities of the Panama Arc (Villamil, 1999; Bermudez et al.,
(Figure 14C), and erosion of previous Paleocene and 2010; Vargas and Mann, 2013).
Cretaceous deposits occurred with more intensity in Pre-Cretaceous depocenters filled with continental
the western CRB than in the eastern CRB (Figure 10D rift facies have been previously documented along the
and E); (4) subsidence during the early Miocene is length of the PR (Miller, 1962; Forero, 1970; Caceres et
related to local flexure produced by uplifting northern al., 1980; Maze, 1984; Ujueta and Llinas, 1990). Simi-
Andean ranges to the west and east of the CRB; this lar rift facies are only present locally southeast of the
flexural event created accommodation space that was SSMM (Gomez Tapias et al., 1969), which suggests that
rapidly filled by deposition of a continental early to the main pre-Cretaceous rift system was parallel to the
middle Miocene units (Figure 14D); and (5) a second present-day location of the PR (Lugo and Mann, 1995;
exhumation event by late Miocene–Pliocene was more Mojica and Kammer, 1995). Our gravity and mag-
severe in the eastern CRB (Figure 10B and G); this netic modeling shows a thick sedimentary to meta-­
rapid exhumation exhumed Jurassic and Cretaceous sedimentary package, especially in the area of the
rocks were previously buried to a depth of 4 km (2.5 northern PR, and is supportive of the presence of the
mi) (Hernandez and Jaramillo, 2009). main Mesozoic rift in this area (Figures 5B, C, and 6F).
The Miocene convergent phase continued into the
Pliocene in the eastern PR (Kellogg, 1981), whereas the
western and central part of the CRB does not show a Implications for Petroleum Systems
conspicuous Pliocene deformation (Figure 10C, D, and
E). Such a pattern is similar to the thermal history of the Past hydrocarbon exploration of the CRB dating back
eastern SSMM, where rapid exhumation occurred dur- to the early 20th century identified the presence of oil
ing the Eocene–Oligocene (Villagomez et al., 2011). The shows, non-commercial oil and gas accumulations, and
Miocene–Pliocene convergent and uplift event—that is one commercial gas field (Garcia et al., 2008). Hydro-
commonly called the “Andean orogeny”—may have carbon discoveries in the Paleocene Hato Nuevo For-
been activated by the collision of the Panama Arc in mation and in Cretaceous formations—including the El
the late Miocene about 12 Ma (Vargas and Mann, 2013) Molino, La Luna, and Cogollo—all support the exist-
­(Figure 12). A very recent (Pleistocene–­Holocene) period ence of source rock maturity, hydrocarbon generation,
of continental deposition has filled most of the CRB, and expulsion (Ayala, 2009) (­ Figure 3). Our basin model
­especially to the west with its flat basin, morphology indicates that the main pulse of hydrocarbon genera-
(Figure 2). Localized active faulting along the eastern tion occurred during a period of rapid Paleocene sub-
CRB has elevated the PR foothills ­(Figure 9). sidence (Figures 10 and 11). Structural mapping shows

13880_ch16_ptg01_431-470.indd 464 10/21/15 11:28 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 465

that pods of mature source rocks are concentrated in regional tectonic events that have affected this area of
structural lows both in the southern CB (La Loma syn- northwestern South America (Figure 1). The main con-
cline) and in the northern RB (Figure 8A and B). The clusions of this study include the following:
location of these productive lows relative to known TOC
hydrocarbon shows and producing areas within the 1. The CRB is underlain by a Precambrian-­Paleozoic
restricted width of the CRB indicates relatively short- continental crust with an average thickness of
distance and mainly vertical migration pathways. 25–30 km (15.5–18.6 mi) and a depth to top crys-
Cretaceous carbonate and thin sandstone reser- talline basement of 0.2–7 km (0.1–4.3 mi) for the
voirs with secondary fracture porosity are the main RB and 7–10 km (4.3–6.2 mi) for the CB ­(Figure 6).
drilling target in the CRB (Mora and Garcia, 2006). Northwest–southeast-­d irected shortening for
Potential structural traps are associated with hang- the basement and overlying sedimentary cover
ing wall closures above major thrusts in the CB is accomplished by high-angle reverse faults that
(Figure 8A and B). Potential seals for hydrocarbon mainly dip southeastward at values of 35–60°
accumulations include the shaly section at the top of ­( Figure 9). The structure of the northern PR is
the late Cretaceous Molino Formation and the silty/ controlled by the late Cenozoic inversion of an
shaly inter-­l ayered coal-bearing Paleogene section ­underlying, NE–SW-trending ­Mesozoic rift basin
(Figure 3). Traps may have been formed during the as proposed by previous workers (­ Figure 6C, D,
two major Cenozoic deformational events that we and F).
have described in this chapter (Figure 12). These 2. The basement-involved structural style of the
deformational events that spanned the Cenozoic region (Figures 4B, 9, and 13) includes high-angle
may have disrupted and isolated the mature source reverse faults that shorten and expose Jurassic to
rock pods as the result of exhumation and erosion Early Cretaceous rocks in the CRB and Paleozoic
that elevated the source rocks above the oil/gas rocks in the PR (Figure 2B). Minor thrust detach-
window. ment levels are controlled by horizons of shaly
The supergiant Maracaibo hydrocarbon basin to Cretaceous rocks (Figure 7).
the east of CRB in Venezuela (Figure 1B) was sepa- 3. Total shortening across the CRB is relatively small:
rated from the CRB by the uplift of the PR in the 6%, 8%, and 6% for the northern, ­central, and south-
Eocene (Figure 12). The Maracaibo Basin experienced ern restored sections, respectively ­(Figure 12). Lim-
two main pulses of hydrocarbon expulsion during ited shortening is much less than that proposed by
the Paleocene and Miocene (Escalona and Mann, Kellogg and Bonini (1982) who postulated lower-
2006b). The first pulse of expulsion is related to the dipping thrust faults in the dip range of 15–20° for
foreland basin produced by the Paleogene oblique a total, NW–SE shortening of 46 km [28.6 mi]) or
collision between the Great Arc of the Caribbean by Montes et al. (2010) who postulated a total of
and the South American plate whereas the second 50–56 km of east-west shortening in the CRB and
pulse is associated with the main late Miocene and PR produced by a ­Miocene to recent, 23–30° clock-
younger uplift of the PR and Venezuelan Andes, or wise rotation by of the Santa Marta Block.
“Andean ­Orogeny” (Escalona and Mann, 2006b). The 4. The CRB is deformed by two distinctive fault
CRB contains fewer and less extensive, high-quality, families: one fault family of high-angle reverse
clastic r­ eservoirs in comparison to the Maracaibo faults that strikes parallel to the long axis of the
Basin because the CRB experienced a more protracted PR (ca N30° E) and another fault family of reverse
and larger magnitude deformational events dur- faults that strikes in a more east–west direction.
ing the Eocene through Miocene than the Maracaibo This second fault family is parallel to a belt of line-
Basin (Figure 4B, 13). The higher level of deforma- aments in the crystalline basement rocks of the
tion and uplift of the CRB relative to the Maracaibo SSMM (Figure 8A and B).
Basin likely reflects the more proximal location of 5. Structure in the northern RB consists of an east-dip-
the CRB to the main C ­ aribbean-South America plate ping monocline disrupted by local reverse faults
boundary. the southern CB shows a more complex configu-
ration where high-angle, east-dipping thrust faults
and local backthrusts control structural highs. Two
Conclusions main lows contain composite fold axes (Figure 8A
and B) to the south, suggesting multi-phase defor-
We have integrated several different types of data mation or the influence of a transpressive regime
to evaluate the crustal structure and deformational associated with the Santa Marta–Bucaramanga
and sedimentary history of the CRB in the context of strike-slip fault system.

13880_ch16_ptg01_431-470.indd 465 10/21/15 11:28 AM


466 Sanchez and Mann

6. Deformational events in the CRB suggest an east- Allen, P. A., and J. R. Allen, 2013, Basin analysis: Principles
ern migration of the deformation as the pattern of and application to petroleum play assessment: Oxford,
preserved depocenters and faulting indicates that John Wiley & Sons, 619 p.
a major uplift occurred first in the west during the Arango, F., and A. Blandon, 2006, Carbones y Lutitas Car-
TOC
bonosas como Rocas Fuentes de Hidrocarburos Terci-
Paleocene–Eocene but then migrated to the east
arios en la Cuenca Cesar-Ranchería Colombia: Evidencias
during late Miocene (Figures 11 and 12).
Petrológicas y Geoquímicas, in Memoir, IX Simposio
7. The CRB resulted from the basin partitioning of ­Bolivariano de Exploración Petrolera en las Cuencas Sub-
the western Marabaibo Basin and uplift of the PR, andinas: Cartagena, Colombia, Asociación Colombiana
whose onset took place by Eocene (Kellogg, 1984; de Geólogos y Geofísicos del Petróleo.
Bayona et al., 2007) (Figure 14B and C). Because of Audemard, F., 1991, Tectonics of western Venezuela: Un-
the proximity of the CRB to the Caribbean arc-con- published Ph.D. dissertation, Rice University, Houston,
tinental collision during the Late Cretaceous–early Texas, 245 p.
Eocene (Cardona et al., 2011a; Escalona and Mann, Audemard, F. E., and I. Serrano, 2001, Future petroliferous
2011; Ayala et al., 2012) and because of the proximity provinces of Venezuela, in M.W. Downey, J.C. Threet, and
of the CRB to the shallowly subducting ­Caribbean W.A. Morgan, eds., Petroleum provinces of the twenty-
first century: AAPG Memoir 74, p. 353–372.
plate (Bernal et al., 2015a), the CRB underwent more
Audemard, F. E., and F. A. Audemard, 2002, Structure of
intensive deformation and protracted uplift in com-
the Merida Andes, Venezuela: Relations with the South
parison to the more distant and sheltered Maracaibo America-Caribbean geodynamic interaction: Tectono-
Basin to the west. physics, v. 345, p. 299–327.
8. Our basin modeling results confirm an effective Ayala, R. C., 2009, Análisis tectonoestratigráfico y de pro-
petroleum system in the southern CRB based cedencia en la subcuenca de Cesar: Relación con los Sis-
on maturation and hydrocarbon generation temas Petroleros: Unpublished M.Sc. thesis, Universidad
from high-quality, Late Cretaceous source rocks SimónBolívar, Caracas, Venezuela, 188 p.
­(Figures 10 and 11). More intensive deformation Ayala, R. C., G. Bayona, A. Cardona, C. Ojeda, O. C. Monte-
and related uplift events adversely affected the negro, C. Montes, V. Valencia, and C. Jaramillo, 2012, The
hydrocarbon potential of the CRB in comparison Paleogene synorogenic succession in the northwestern
Maracaibo Block: Tracking intraplate uplifts and changes
to the less deformed and eroded Maracaibo Basin
in sediment delivery systems: Journal of South American
located to the east of the CRB.
Earth Sciences, v. 39, p. 93–111.
Baranov, V., and H. Naudy, 1964, Numerical calculation of
the formula of reduction to the magnetic pole: Geophys-
Acknowledgments ics, v. 29, p. 67–79.
Bayona, G., F. Lamus-Ochoa, A. Cardona, C. A. Jaramillo,
Seismic information and well data that formed the C. Montes, and N. Tchegliakova, 2007, Provenance ­analysis
basis for this study were kindly provided by the Agen- and Paleocene orogenic processes in the Rancheria Basin
cia Nacional de Hidrocarburos (ANH) of Colombia, (Guajira, Colombia) and surrounding areas: Geologia
and Elsa Jaimes at OGX. We would like to thank Drs. Colombiana, v. 32, p. 21–46.
Bayona, G., G. Jiménez, C. Silva, A. Cardona, C. Montes,
Michael Murphy, Kevin Burke, and Alejandro Escalona
J. Roncancio, and U. Cordani, 2010, Paleomagnetic data
for discussions of the structural interpretation and
and K–Ar ages from Mesozoic units of the Santa Marta
regional tectonics of the area and Dr. Dale Bird for dis- massif: A preliminary interpretation for block ­rotation
cussions of gravity and magnetic interpretations and and translations: Journal of South American Earth
for arranging our access to Oasis Montaj software. Spe- Sciences, v. 29, p. 817–831.
cial thanks to the industry sponsors of the Caribbean Bayona, G., C. Montes, A. Cardona, C. Jaramillo, G. Ojeda,
Basins, Tectonics, and Hydrocarbons Project (CBTH) V. Valencia, and C. Ayala-Calvo, 2011, Intraplate subsidence
for their continued support of our studies. Construc- and basin filling adjacent to an oceanic arc–continent
tive reviews by Georgina Guzman and Bonnie Milne collision: A case from the southern Caribbean‐South
greatly improved the quality of the manuscript. America plate margin: Basin Research, v. 23, p. 403–422.
Bermudez, M. A., B. P. Kohn, P. A. van der Beek, M. Bernet,
P. B. O’Sullivan, and R. Shagam, 2010, Spatial and ­temporal
patterns of exhumation across the Venezuelan ­Andes:
References Cited Implications for Cenozoic Caribbean geodynamics:
Tectonics, v. 29, TC5009 p., doi: 10.1029/2009TC002635
Alberding, H., 1957, Application of principles of wrench- Bermudez, M. A., P. Van der Beek, and M. Bernet, 2011,
fault tectonics of Moody and Hill to northern South Asynchronous Miocene–Pliocene exhumation of the cen-
America: GSA Bulletin, v. 68, p. 785–790. tral Venezuelan Andes: Geology, v. 39, p. 139–142.

13880_ch16_ptg01_431-470.indd 466 10/21/15 11:28 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 467

Bernal-Olaya, R., P. Mann, and C. A. Vargas, 2015a, Earth- Cardona Molina, A., U. G. Cordani, and W. D. MacDonald,
quake, tomography, seismic reflection and gravity 2006, Tectonic correlations of pre-Mesozoic crust from the
­e vidence for a shallowly-dipping subduction zone northern termination of the Colombian Andes, Caribbean
­beneath the Caribbean Margin of Colombia, in C. B
­ artolini, region: Journal of South American Earth Sciences, v. 21,
TOC
V. ­Ramirez, L. Ardila, and M. Ramirez, eds., P
­ etroleum p. 337–354.
­g eology and hydrocarbon potential of ­C olombia Case, J. E., and W. R. MacDonald, 1973, Regional gravity
­Caribbean Margin: AAPG Memoir 108, p. 247–270. anomalies and crustal structure in northern Colombia:
Bernal-Olaya, R., J. Sanchez, P. Mann, and M. Murphy, 2015b, GSA Bulletin, v. 84, p. 2905–2916.
Along-strike crustal thickness variations of the subduct- Castillo, V., 2001, Structural analysis of Cenozoic fault sys-
ing Caribbean plate produces two distinctive styles of tems using 3-D seismic reflection data in the s­ outhern
thrusting in the offshore South Caribbean deformed Maracaibo Basin: Unpublished Ph.D. dissertation,
belt, Colombia, in C. Bartolini, V. Ramirez, L. Ardila, and Austin, Texas, University of Texas, 188 p.
M. Ramirez, eds., Petroleum geology and hydrocarbon Cediel, F., R. P. Shaw, and C. Caceres, 2003, Tectonic assem-
­potential of Colombia Caribbean Margin: AAPG Memoir bly of the northern Andean Block, in C. Bartolini, R. T.
108, p. 295–322. Buffer, and J. Blickwede, eds., The Circum-Gulf of Mexico
Bernal-Olaya, P. Mann, and A. Escalona, 2015c, Tecton- and the Caribbean: Hydrocarbon habitats, basin forma-
ostratigraphic evolution of the Lower Magdalena Basin, tion, and plate tectonics: AAPG Memoir 79, p. 815–848.
Colombia: An example of an underfilled to overfilled Ceron-Abril, J. F., 2008, Crustal structure of the Colombian
forearc basin, in C. Bartolini, V. Ramirez, L. Ardila, and Caribbean Basin and margins: Unpublished Ph.D. dis-
M. Ramirez, eds., Petroleum geology and hydrocarbon sertation, Columbia, South Carolina, University of South
potential of Colombia Caribbean Margin: AAPG Memoir Carolina., 165 p.
108, p. 345–398. Colletta, B., F. Roure, B. Toni, D. Loureiro, H. Passalacqua,
Blakely, R. J., 1996, Potential theory in gravity and magnetic and Y. Gou, 1997, Tectonic inheritance, crustal architec-
applications: Cambridge, Cambridge University Press, ture, and contrasting structural styles in the Venezuela
464 p. Andes: Tectonics, v. 16, p. 777–794.
Burke, K., 1988, Tectonic evolution of the Caribbean: Annual Colmenares, L., and M. D. Zoback, 2003, Stress field and
Review of Earth and Planetary Sciences, v. 16, p. 201–230. seismotectonics of northern South America: Geology,
Caceres, H., R. Camacho, and J. Reyes, 1980, Guidebook to v. 31, p. 721–724.
the geology of the Rancheria Basin, in Geotec ed., Geolog- Cordani, U. G., A. Cardona, D. M. Jimenez, D. Liu, and A.
ical field-trips, Colombia 1980–198, Asociacion Colombi- P. Nutman, 2005, Geochronology of Proterozoic basement
ana de Geologos y Geofísicos del Petroleo, p. 1–31. inliers in the Colombian Andes: Tectonic history of rem-
Calais, E., and P. Mann, 2009, A combined GPS velocity field nants of a fragmented Grenville Belt, in A. P. M. Vaughan,
for Caribbean plate and its margins, EOS, transactions of P. T. Leat, and R. J. Pankhurst, eds., Terrane processes at
the American Geophysical Union, AGU fall meeting, San the margins of Gondwana: GSL Special Publication 246,
Francisco, California, v. 1, 0657 p. p. 329–346.
Campbell, C. J., 1965, The Santa Marta wrench fault of Cortes, M., B. Colletta, and J. Angelier, 2006, Structure and
­Colombia and its regional setting, in J.B. Saunders, ed., tectonics of the central segment of the eastern Cordillera
Transactions of the Fourth Caribbean Geological Confer- of Colombia: Journal of South American Earth Sciences,
ence: Port-of-Spain, Trinidad, v. 4, p. 247–261. v. 21, p. 437–465.
Cardona, A., D. Chew, V. A. Valencia, G. Bayona, A. Misko- Dahlstrom, C., 1969, Balanced cross sections: Canadian Jour-
vic, and M. Ibanez-Mejia, 2010, Grenvillian remnants in nal of Earth Sciences, v. 6, p. 743–757.
the Northern Andes: Rodinian and Phanerozoic paleoge- Dengo, C. A., and M. C. Covey, 1993, Structure of the eastern
ographic perspectives: Journal of South American Earth Cordillera of Colombia; implications for trap styles and
Sciences, v. 29, p. 92–104. regional tectonics: AAPG Bulletin, v. 77, p. 1315–1337.
Cardona, A., V. A. Valencia, G. Bayona, J. Duque, M. Ducea, Duerto, L., A. Escalona, and P. Mann, 2006, Deep struc-
G. Gehrels, C. Jaramillo, C. Montes, G. Ojeda, and J. Ruiz, ture of the Merida Andes and Sierra de Perija mountain
2011a, Early-subduction-related orogeny in the northern fronts, Maracaibo Basin, Venezuela: AAPG Bulletin, v. 90,
Andes: Turonian to Eocene magmatic and provenance p. 505–528.
record in the Santa Marta massif and Rancheria Basin, De Toni, B., and J. N. Kellogg, 1993, Seismic evidence for
northern Colombia: Terra Nova, v. 23, p. 26–34. blind thrusting of the northwestern flank of the Venezue-
Cardona, A., V. Valencia, M. Weber, J. Duque, C. Montes, lan Andes: Tectonics, v. 12, p. 1393–1409.
G. Ojeda, P. Reiners, K. Domanik, S. Nicolescu, and D. Duque-Caro, H., 1979, Major structural elements and evo-
Villagomez, 2011b, Transient Cenozoic tectonic stages lution of northwestern Colombia, in J. S. Watkins, L.
in the southern margin of the Caribbean plate: U-Th/ Montadert, and P. W. Dickerson, eds., Geological and
He thermochronological constraints from Eocene geophysical investigations of continental margins: AAPG
­p lutonic rocks in the Santa Marta massif and Serranía Memoir 29: p. 329–351.
de Jarara, northern Colombia: Geologica Acta, v. 9, Duque-Caro, H., 1990, Neogene stratigraphy, paleocean-
p. 445–466. ography and paleobiogeography in northwest South

13880_ch16_ptg01_431-470.indd 467 10/21/15 11:28 AM


468 Sanchez and Mann

America and the evolution of the Panama Seaway: Gose, W., A. Perarnau, and J. Castillo, 2003, Paleomagnetic
­Palaeogeography, Palaeoclimatology, Palaeoecology, v. results from the Perija mountains, Venezuela: An exam-
77, p. 203–234. ple of vertical axis rotation, in C. Bartolini, R. Buffler, and
Escalona, A., and P. Mann, 2006a, Tectonic controls of the J. Blickwede, eds., The Circum-Gulf of Mexico and the
TOC
right-lateral Burro Negro tear fault on Paleogene struc- Caribbean: Hydrocarbon habitats, basin formation, and
ture and stratigraphy, northeastern Maracaibo Basin: plate tectonics AAPG Memoir 79, p. 177–178.
AAPG Bulletin, v. 90, p. 479–504. Gulick, S. P. S., N. L. B. Bangs, G. F. Moore, J. Ashi, K. M.
Escalona, A., and P. Mann, 2006b, An overview of the petroleum Martin, D. S. Sawyer, H. J. Tobin, S. Kuramoto, and A.
system of Maracaibo Basin: AAPG Bulletin, v. 90, p. 657–678. Taira, 2010, Rapid forearc basin uplift and megasplay
Escalona, A., and P. Mann, 2011, Tectonics, basin subsidence fault development from 3-D seismic images of Nankai
mechanisms, and paleogeography of the ­Caribbean-South Margin off Kii Peninsula, Japan: Earth and Planetary
American plate boundary zone: Marine and Petroleum Science Letters, v. 300, p. 55–62.
Geology, v. 28, p. 8–39. Hernandez, O., and J. M. Jaramillo, 2009, Reconstruccion
Eva, A. N., K. Burke, P. Mann, and G. Wadge, 1989, Four- de la historia termal en los sectores de Luruaco y Cerro
phase tectonostratigraphic development of the southern Cansona-cuenca del Sinu-San Jacinto y el piedemonte oc-
Caribbean: Marine and Petroleum Geology, v. 6, p. 9–21. cidental de la Serrania del Perija entre Codazzi y la Jagua
Ewing, J. I., J. Antoine, and M. Ewing, 1960, Geophysical de Ibirico-cuenca Cesar-Rancheria, Agencia Nacional de
­measurements in the western Caribbean and Gulf of M ­ exico: Hidrocarburos: Bogota, Colombia, 58 p.
Journal of Geophysical Research, v. 65, p. 4087–4126. Jacobsen, B., 1987, A case for upward continuation as a
Feo-Codecido, G., F. D. J. Smith, N. Aboud, and E. Di standard separation filter for potential‐field maps:
­G iacomo, 1984, Basement and Paleozoic rocks of the Geophysics, v. 52, p. 1138–1148.
­Venezuelan Llanos Basins, in R. B. Hargraves, R. Shagam, Kammer, A., and J. Sanchez, 2006, Early Jurassic rift struc-
and W. E. Bonini, eds., The Caribbean-South American tures associated with the Soapaga and Boyacá faults of
plate boundary and regional tectonics: GSA Memoir 162: the eastern Cordillera, Colombia: Sedimentological infer-
p. 175–187. ences and regional implications: Journal of South Ameri-
Feuillet, N., I. Manighetti, and P. Tapponnier, 2002, Arc par- can Earth Sciences, v. 21, p. 412–422.
allel extension and localization of volcanic complexes Kellogg, J. N., 1981, Cenozoic basement tectonics of the
in Guadeloupe, Lesser Antilles: Journal of Geophysical Sierra de Perija, Venezuela and Colombia: Unpublished
Research, Solid Earth (1978–2012), v. 107, no. B12 (2002), Ph.D. dissertation, Princeton, New Jersey, Princeton
ETG-3. ­University, 236 p.
Florez-Nino, J., 2001, Elastic geomechanical model of Bucar- Kellogg, J. N., and W. E. Bonini, 1982, Subduction of the Car-
amanga and Oca faults and the origin of the Sierra Ne- ibbean plate and basement uplifts in the overriding South
vada de Santa Marta, Northern Andes, Colombia, EOS, American plate: Tectonics, v. 1, p. 251–276.
transactions of the American Geophysical Union, AGU Kellogg, J., 1984, Cenozoic tectonic history of the Sierra de
fall meeting, San Francisco, California, v. 1, p. 0834. Perija, Venezuela-Colombia, and adjacent basins, in W.E.
Forero, A., 1970, Estartigrafia del Pre-Cretacico en el flanco Bonini, R.B. Hargraves, and R. Shagam, eds., The Carib-
occidental de la Serrania de Perija: Geologia Colombiana, bean-South American plate boundary and regional tecton-
v. 7, p. 7–78. ics: GSA Memoir 162, p. 239–261.
Gallango, O., E. Novoa, and A. Bernal, 2002, The petroleum Kennan, L., and J. Pindell, 2009, Dextral shear, terrane ac-
system of the central Perija fold belt, western Venezuela: cretion and basin formation in Northern Andes: Best
AAPG Bulletin, v. 86, p. 1263–1284. explained by interaction with Pacific-derived Caribbean
Garcia, M., R. Umana, A. Arias, Y. Cortes, M. Moreno, O. Plate?, in K. James, M. A. Lorente, and J. Pindell, eds., The
­Salazar, and M. Jimenez, 2008, Prospectividad de la Cuenca geology and evolution of the region between North and
Cesar Rancheria: Bucaramanga, Universidad Industrial de South America: GSL, Special Publication 328, p. 487–531.
Santander, Agencia Nacional de Hidrocarburos, 244 p. Kroehler, M.E., P. Mann, A. Escalona, and G.L. Christeson,
Gomez, E., T. E. Jordan, R. W. Allmendinger, and N. Cardozo, 2011, Late Cretaceous-Miocene diachronous onset of
2005a, Development of the Colombian foreland-­b asin back thrusting along the South Caribbean deformed belt
system as a consequence of diachronous exhumation of and its importance for understanding processes of arc
the northern Andes: GSA Bulletin, v. 117, p. 1272–1292. collision and crustal growth: Tectonics, v. 30, TC6003,
Gomez, E., T. E. Jordan, R. W. Allmendinger, K. Hegarty, and doi:10.1029/2011TC002918.
S. Kelley, 2005b, Syntectonic Cenozoic sedimentation in Lugo, J., and P. Mann, 1995, Jurassic-Eocene tectonic evo-
the northern middle Magdalena Valley Basin of Colombia lution of Maracaibo Basin, Venezuela, in A. J. Tankard,
and implications for exhumation of the Northern Andes: R. Suarez S., and H. J. Welsink, eds., Petroleum basins of
GSA Bulletin, v. 117, p. 547–569. South America: AAPG Memoirs 62, p. 699–725.
Gomez Tapias, J., A. Nivia Guevara, N. E. Montes Ramierez, MacDonald, W.D., and N.D. Opdyke, 1984, Preliminary
D. M. Jimenez Mejia, M. L. Tejada Avella, M. J. Sepulveda paleomagnetic results from Jurassic rocks of the Santa
Ospina, J. A. Osorio Naranjo, T. Gaona Narvaez, H. Died- Marta massif, Colombia: GSA Memoir, v. 162, p. 295–298.
erix, H. Uribe Pena, and M. Mora Penagos, 2007, Mapa Mann, P., and K. Burke, 1984, Neotectonics of the Carib-
Geologico de Colombia: Bogota, Colombia, Instituto Co- bean: Review of Geophysics and Space Physics, v. 22,
lombiano de Geologia y Mineria. p. 309–362.

13880_ch16_ptg01_431-470.indd 468 10/21/15 11:28 AM


Integrated Structural and Basinal Analysis of the Cesar–Rancheria Basin, Colombia 469

Mann, P., A. Escalona, M.V. Castillo, 2006, Regional geologic ­ olombia interpreted from fission track results and struc-
C
and tectonic setting of the Maracaibo supergiant basin, tural relationships: Implications for petroleum systems:
western Venezuela: AAPG Bulletin, v. 90, p. 445-478. AAPG Bulletin, v. 94, p. 1545–1580.
Mauffret, A., and S. Leroy, 1997, Seismic stratigraphy and Mora, A., A. Reyes-Harker, G. Rodriguez, E. Teson, J. C.
TOC
structure of the Caribbean igneous province: Tectono- Ramirez-Arias, M. Parra, V. Caballero, J. P. Mora, I.
physics, v. 283, p. 1–4. Quintero, V. Valencia, M. Ibanez, B. K. Horton, and D. F.
Maus, S., U. Barckhausen, H. Berkenbosch, N. Bournas, Stockli, 2013, Inversion tectonics under increasing rates of
J. Brozena, V. Childers, F. Dostaler, J. D. Fairhead, C. shortening and sedimentation: Cenozoic example from
Finn, R. R. B. von Frese, C. Gaina, S. Golynsky, R. Kucks, the Eastern Cordillera of Colombia, in Nemcok, M., A.
H. Luhr, P. Milligan, S. Mogren, D. Muller, O. Olesen, Mora, and J.W. Cosgrove, eds., Thick-skin-dominated
M. Pilkington, R. Saltus, B. Schreckenberger, E. Thebault, orogens: From initial inversion to full accretion: GSL,
and F. Caratori Tontini, 2009, EMAG2: A 2-arc-minute Special Publications, v. 377, p. 411–442.
resolution Earth Magnetic Anomaly Grid compiled from Nettleton, L. L., 1971, Gravity and magnetics for geologists and
satellite, airborne and marine magnetic measurements: seismologists: Society of Exploration Geophysicists. 121 p.
Geochemistry, Geophysics, Geosystems, v. 10, no. 8, p. 1–12 Nova, G., P. Montano, G. Bayona, A. Rapalini, and
Maze, W. B., 1984, Jurassic La Quinta Formation in the Sierra C. Montes, 2012, Paleomagnetismo en rocas del Jurasico y
de Perijá, northwestern Venezuela: Geology and tectonic Cretacico inferior en el flanco occidental de la serrania del
environment of red beds and volcanic rocks, in W. E. Perija; Contribuciones a la evolucion tectonica del NW de
Bonini, R. B. Hargraves, and R. Shagam, eds., The Carib- Suramerica: Boletín de Geologia, v. 34, p. 117–138.
bean-South American plate boundary and regional tecton- Parnaud, Y., Y. Gou, J. Pascual, M. A. Capello, I. Truskowski,
ics: GSA Memoir 162, p. 263–282. and H. Passalacqua, 1995, Stratigraphic synthesis of
McKenzie, D., 1978, Some remarks on the development of Western Venezuela, in A. J. Tankard, R. Suarez S., and
sedimentary basins: Earth and Planetary Science Letters, H. J. Welsink, eds., Petroleum basins of South America:
v. 40, p. 25–32. AAPG Memoir 62, p. 681–698.
Miller, J. B., 1962, Tectonic trends in the Sierra de Perija Parra, M., A. Mora, C. Lopez, L. E. Rojas, and B. K. ­Horton,
and adjacent parts of Venezuela and Colombia: AAPG 2012, Detecting earliest shortening and deformation
­Bulletin, v. 46, p. 1565–1595. advance in thrust-belt hinterlands: Example from the
Mojica, J., and A. Kammer, 1995, Eventos Jurasicos de ­Colombian Andes: Geology, v. 40, p. 175–178.
­Colombia: Geologia Colombiana, v. 19, p. 165–172. Parra, M., A. Mora, E. R. Sobel, M. R. Strecker, and R.
Montes, C., R. D. Hatcher Jr., and P. A. Restrepo-Pace, 2005a, ­G onzalez, 2009, Episodic orogenic-front migration in
Tectonic reconstruction of the northern Andean blocks: the northern Andes: Constraints from low-temperature
Oblique convergence and rotations derived from the kin- ­thermochronology in the Eastern Cordillera, Colombia:
ematics of the Piedras-Girardot area, Colombia: Tectono- Tectonics, v. 28, TC4004, doi:10.1029/2008TC002423.
physics, v. 399, p. 221–250. Pindell, J., and K. Tabbutt, 1995, Mesozoic – Cenozoic
Montes, C., G. Bayona, C. Jaramillo, G. Ojeda, M. ­Molina, ­A ndean paleogeography and regional controls on
and F. Herrera, 2005b, Uplift of the Sierra Nevada de ­hydrocarbon systems, in A. J. Tankard, R. Suarez S., and
Santa Marta and subsidence in the Cesar-Rancheria H. J. Welsink, eds., Petroleum basins of South America:
­Valley: Rigid-beam pivot model, in Extended abstracts, AAPG Memoir 62, p. 101–128.
Sixth International Symposium of Andean Geodynamics, Ramsay, J., and M. Huber, 1983, Strain analysis, the
­Barcelona, Spain, p. 520–523. ­t echniques of modern structural geology 1: London,
Montes, C., G. Guzman, G. Bayona, A. Cardona, V. Valencia, ­Academic Press, 308 p.
and C. Jaramillo, 2010, Clockwise rotation of the Santa Restrepo-Pace, P. A., J. Ruiz, G. Gehrels, and M. Cosca, 1997,
Marta massif and simultaneous Paleogene to Neogene Geochronology and Nd isotopic data of G ­ renville-age
deformation of the Plato-San Jorge and Cesar-Rancheria rocks in the Colombian Andes: New constraints for
Basins: Journal of South American Earth Sciences, v. 29, late Proterozoic-early Paleozoic paleocontinental
p. 832–848. ­reconstructions of the Americas: Earth and Planetary
Mora, A., and A. Garcia, 2006, Cenozoic tectono-strati- ­Science ­Letters, v. 150, p. 427–441.
graphic relationships between the Cesar Subbasin and Reyes-Santos, J., M. Mantilla, and J. S. Gonzalez, 2000, Re-
the southeastern Lower Magdalena Valley Basin of north- giones tectono-sedimentarias del Valle inferior del
ern Colombia: Search and Discovery Article #30046 (2006) Magdalena, Colombia, in Memorias VII Simposio Bo-
Posted November 27, 2006. *Adapted from poster presen- livariano, Exploración Petrolera en las Cuencas Sub-
tation at AAPG 2006 Annual Convention and Exhibition, andinas: Caracas, Venezuela, Sociedad Venezolana de
Houston, Texas. Geólogos, p. 310–333.
Mora, A., M. Parra, M. R. Strecker, E. R. Sobel, H. Hooghiem- Rincon, D. A., J. E. Arenas, C. H. Cuartas, A. L. Cardenas,
stra, V. Torres, and J. Vallejo-Jaramillo, 2008, Climatic C. E. Molinares, C. Caicedo, and C. A. Jaramillo, 2007,
forcing of asymmetric orogenic evolution in the eastern Eocene-Pliocene planktonic forminifera biostratigraphy
Cordillera of Colombia: GSA Bulletin, v. 120, p. 930–949. from the continental margin of the southwest Caribbean:
Mora, A., B. Horton, A. Mesa, J. Rubiano, R. Ketcham, M. Stratigraphy, v. 4, p. 261–312.
Parra, V. Blanco, D. Garcia, and D. Stockli, 2010, Migra- Roeder, D., and R. Chamberlain, 1995, Eastern Cordillera
tion of Cenozoic deformation in the eastern C­ ordillera of of Colombia: Jurassic-Neogene crustal evolution, in

13880_ch16_ptg01_431-470.indd 469 10/21/15 11:28 AM


470 Sanchez and Mann

A. J. Tankard, R. Suarez S., and H. J. Welsink, eds., Marta y la Falla de Bucaramanga: Geologia Colombiana,
­Petroleum basins of South America: AAPG Memoir 62, p. v. 28, p. 133–153.
633–645. Ujueta, G., and R. Llinas, 1990, Reconocimiento geologico
Sanchez, J., B. K. Horton, E. Teson, A. Mora, R. A. Ket- de la parte septentrional de la Sierra de Perija: Geologia
TOC
cham, and D. F. Stockli, 2012, Kinematic evolution of ­Colombiana, v. 17, p. 197–209.
Andean fold‐thrust structures along the boundary be- Van der Hilst, R., and P. Mann, 1994, Tectonic implications
tween the Eastern Cordillera and Middle Magdalena of tomographic images of subducted lithosphere beneath
Valley Basin, Colombia: Tectonics, v. 31, TC3008, northwestern South America: Geology, v. 22, p. 451–454.
doi:10.1029/2011TC003089. Vargas, C. A., and P. Mann, 2013, Tearing and breaking off
Sandwell, D., and W. Smith, 1997, Marine gravity anomaly of subducted slabs as the result of collision of the Pan-
from Geosat and ERS 1 satellite altimetry: Journal of ama Arc-indenter with northwestern South America:
­Geophysical Research B, v. 102, p. 10039–10054. Bulletin of the Seismological Society of America, v. 103,
Shagam, R., B. P. Kohn, P. O. Banks, L. E. Dasch, R. Vargas, p. 2025–2046.
G. I. Rodriguez, and N. Pimentel, 1984, Tectonic impli- Vence, E., 2008, Subsurface structure, stratigraphy, and
cations of Cretaceous–Pliocene fission track ages from ­regional tectonic controls of the Guajira Margin of north-
rocks of the Circum-Maracaibo Basin region of western ern Colombia: Unpublished M.Sc. thesis, Austin, Texas,
Venezuela and eastern Colombia, in W. E. Bonini, R. B. University of Texas, 128 p.
Hargraves, and R. Shagam, eds., The Caribbean-South Verduzco, B., J. Fairhead, C. Green, and C. MacKenzie, 2004,
American plate boundary and regional tectonics: GSA New insights into magnetic derivatives for structural
Memoir 162, p. 385–412. mapping: The Leading Edge, v. 23, p. 116–119.
Taboada, A., L. A. Rivera, A. Fuenzalida, A. Cisternas, Villagomez, D., R. Spikings, A. Mora, A. Cardona, G.
P. Herve, B. Harmen, J. Olaya, and C. Rivera, 2000, ­Guzman, G. Ojeda, E. Cortes, and R. Van der Lelij, 2011,
­Geodynamics of the northern Andes: Subductions and Vertical tectonics at a continental crust-oceanic plateau,
intracontinental deformation (Colombia): Tectonics, v. 19, plate boundary zone: Low temperature thermochronol-
p. 787–813. ogy of the Sierra Nevada de Santa Marta, Colombia:
Teson, E., A. Mora, A. Silva, J. Namson, A. Teixell, J. ­Tectonics, v. 30, TC4004, doi:10.1029/2010TC002835.
­Castellanos, W. Casallas, M. Julivert, M. Taylor, and M. Villamil, T., 1999, Campanian-Miocene tectonostratigraphy,
Ibanez-Mejia, 2013, Relationship of Mesozoic Graben de- depocenter evolution and basin development of Colom-
velopment, stress, shortening magnitude, and structural bia and western Venezuela: Palaeogeography, v. 153,
style in the eastern Cordillera of the Colombian Andes, p. 239–275.
in M. Nemcok, A. Mora, and J.W. Cosgrove, eds., Thick- Wygrala, B. P., 1989, Integrated study of an oil field in the
skin-dominated orogens: From initial inversion to full southern Po Basin, northern Italy: Unpublished Ph.D. dis-
­accretion: GSL, Special Publication 377, p. 257–283. sertation, Cologne, Germany, University of Koln, 324 p.
Tissot, B., R. Pelet, and P. Ungerer, 1987, Thermal history Zeng, H., D. Xu, and H. Tan, 2007, A model study for esti-
of sedimentary basins, maturation indices, and kinet- mating optimum upward-continuation height for grav-
ics of oil and gas generation: AAPG Bulletin, v. 71, ity separation with application to a Bouguer gravity
p. 1445–1466. anomaly over a mineral deposit, Jilin province, northeast
Tissot, B. P., and D. H. Welte, 1984, Petroleum formation and China: Geophysics, v. 72, p. I45–I50.
occurrence: Berlin, Springer-Verlag, 702 p. Zuluaga C., A. Pinilla, P. Mann, 2015, Jurassic silicic
Tschanz, C. M., R. F. Marvin, J. Cruz-B, H. H. Menhert, ­v olcanism and associated continental-arc basin in
and G. T. Cebula, 1974, Geologic evolution of the Sierra northwestern Colombia (southern boundary of the
­Nevada de Santa Marta, northeastern Colombia: GSA ­Caribbean plate), in C. Bartolini, V. Ramirez, L. Ardila,
Bulletin, v. 85, p. 273–284. and M. ­R amirez, eds., Petroleum geology and hydro-
Ujueta, G., 2003, La Falla de Santa Marta-Bucaramanga no es carbon ­potential of Colombia Caribbean Margin: AAPG
una sola falla; son dos fallas diferentes: La Falla de Santa ­Memoir 108, p. 429–468.

13880_ch16_ptg01_431-470.indd 470 10/21/15 11:28 AM

You might also like