You are on page 1of 24

Journal of South American Earth Sciences 121 (2023) 104109

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


journal homepage: www.elsevier.com/locate/jsames

Neogene-Quaternary evolution of the La Baja Guajira basin, Colombia:


Burial and thermal history with implications on petroleum systems
Leidy Castro-Vera a, b, *, Ralf Littke a, Stefan Back a, Rocío Bernal-Olaya c, Martin Reyes b, d,
J. Alejandro Mora-Bohórquez e, A. Johamna Romero e
a
Energy and Mineral Resources Group (EMR), Institute of Geology and Geochemistry of Petroleum and Coal, RWTH Aachen University, Lochnerstraße 4-20, 52056,
Aachen, Germany
b
Grupo de Investigación en Ciencias de la Tierra y Energía, Amonite S.A.S, Bucaramanga, Santander, Colombia
c
Universidad Industrial de Santander, Carrera 27 Calle 9, Bucaramanga, Santander, Colombia
d
Department of Structural Geology and Geodynamics, Georg-August-Universität Göttingen, Göttingen, Germany
e
Hocol S.A., Carrera 7 No. 113-43, 16th Floor, Bogotá, Colombia

A R T I C L E I N F O A B S T R A C T

Keywords: The La Baja Guajira Basin is South America’s northernmost part that extends from onshore Colombia into the
La Baja Guajira basin Caribbean Sea and represents an important gas-producing region in Colombia. The dataset used in this study
Seismic interpretation consists of twenty-four (24) regional 2D seismic-reflection lines integrated with information from nine (9)
Sedimentation
boreholes across the basin. We present seismic interpretations of key stratigraphic and structural features con­
Inversion
Subsidence
strained by 1D basin modeling. This allows us to describe the Neogene-Quaternary evolution of the basin, and
Maturity understand the possible implications for the hydrocarbon potential.
1D modeling We have interpreted eight (8) seismic units within the La Baja Guajira Basin (acoustic basement to U7) affected
by three fault groups: 1) NW-SE-striking basement normal faults, 2) inverted normal faults, and 3) strike-slip
faults. Considering the evidence, we propose four tectonic phases along the basin’s evolution. The basement
was affected by normal faulting as the result of extensional tectonics prior to the deposition of the sedimentary
infill (Paleogene?). Another extensional interval occurred during the Early Miocene, triggering a moderate ac­
commodation within the proximal area. Subsequently, extension migrated westwards resulting in the formation
of the main La Baja Guajira Basin depocenter. In the Middle Miocene (post-extension), sedimentation was
controlled by exhumation pulses in adjacent areas to the basin, which resulted in significant sediment input
contemporaneous with rapid tectonic subsidence. Yet, local inversion of NW-SE-oriented basement-rooted faults
affected today’s offshore domain until the Late Miocene. This inversion can be associated with regional
compressive deformation linked to the activity of the Oca and Cuiza faults. A second episode of rapid tectonic
subsidence happened in the Early Pliocene. In the south of the basin, rocks were subjected to maximum burial
and highest temperatures in recent times, conditions that might lead to hydrocarbon generation. The high
Neogene to recent sedimentation rates calculated for the La Baja Guajira Basin support microbial gas generation;
structural reactivation in this time supported trap formation, particularly in the offshore parts of the basin.

1. Introduction to be part of the Caribbean basins, and to be influenced by the regional-


scale interaction between the Caribbean Arc and the South American
The La Baja Guajira Basin is located within the Guajira Peninsula, continent (e.g. Lugo and Mann, 1995; Pindell et al., 2006; Pindell and
which represents the northernmost portion of the South American Plate Kennan, 2009; Escalona and Mann, 2011).
( Fig. 1). The basin lies between longitudes 72◦ 10′ 6 W and 73◦ 56′ 45 W The La Baja Guajira Basin is the primary gas-producing region of
and latitudes 11◦ 16′ 32 N and 12◦ 06′ 32 N, and comprises an area of Colombia, where hydrocarbon production takes place on offshore plat­
approximately 17,000 km2 (Figs. 1 and 2). The study area is considered forms at Chuchupa (~3.5 Trillion Cubic Feet - TCF) and Ballena (~1.26

* Corresponding author. Energy and Mineral Resources Group (EMR), Institute of Geology and Geochemistry of Petroleum and Coal, RWTH Aachen University,
Lochnerstraße 4-20, 52056, Aachen, Germany.
E-mail address: leidy.castro@emr.rwth-aachen.de (L. Castro-Vera).

https://doi.org/10.1016/j.jsames.2022.104109
Received 29 June 2022; Received in revised form 7 November 2022; Accepted 9 November 2022
Available online 15 November 2022
0895-9811/© 2022 Elsevier Ltd. All rights reserved.
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

TCF), and onshore at the Riohacha Field (~0.09 TCF). The basin rep­ geometry and its Neogene-Quaternary tectonic evolution in a more in­
resents an economically interesting area due to its proximity to large tegrated manner by covering all the domains of the basin (offshore and
hydrocarbon deposits in the Maracaibo Basin, and the recent discoveries onshore), and to evaluate possible implications for the hydrocarbon
of the Orca-1 well in 2014 located offshore La Alta Guajira Basin and the potential.
Uchuva-1 well in 2022 located nearby Well-1 in the offshore study area In this article, we present seismic sections that allowed us to identify
(Fig. 2). regional structural features and their extent within the La Baja Guajira
Cretaceous to Cenozoic tectonic events affecting the Car­ Basin, and stratigraphic relationships. We correlated the results from the
ibbean–South American boundary have previously been reported (i.e. seismic interpretation with burial and thermal histories that resulted
Mann et al., 1990; Pindell and Barrett, 1990; Avé Lallemant, 1997; from 1D basin modeling to constrain the main tectonic phases that
Gorney et al., 2007; Escalona and Mann, 2011; Londoño et al., 2015; affected the study area during the Neogene and Quaternary and evaluate
Vence and Mann, 2020) resulting in the proposal of several models. possible implications for the hydrocarbon potential. The results are
These studies have mainly focused on the tectonic interpretation of the important for the further understanding of the evolution of northwest
La Alta Guajira Basin, and locally on the offshore area of La Baja Guajira Colombia, and for evaluating possible hydrocarbon potential in the area.
Basin, while the onshore areas have remained less integrated. Therefore,
in this study, both onshore and offshore areas of La Baja Guajira Basin
are analyzed. Our main objective is to be able to comprehend the basin’s

Fig. 1. Tectonic map of northwestern South America showing the location of the study area in the north of Colombia (yellow rectangle) and the main structures. The
digital elevation model is taken from the Global Mapper Online Data Source (2020) and bathymetry is taken from the General Bathymetric Chart of Oceans (2020).
The vectors of plate motion are shown in white arrows with the relative rates of movements (Taboada et al., 2000). SNSM= Sierra Nevada de Santa Marta. WC =
Western Cordillera; CC = Central Cordillera; EC = Eastern Cordillera; SCDB = South Caribbean Deformed Belt. The fault traces are shown in the solid black lines, and
labeled as follows RFZ = Romeral Fault Zone, SMBF = Santa Marta – Bucaramanga Fault, CF = Cuiza Fault, OF = Oca Fault, BFZ = Boconó Fault Zone. The dashed
lines represent tectonic provinces proposed by Escalona and Mann (2011).

2
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

2. Geological setting to 20 km of horizontal offset (Feo-Codecido, 1972), 30 km (Soulas et al.,


1987; Audemard, 1993; Audemard et al., 1994) to 60 km (Tschanz et al.,
2.1. Structural framework 1974) of right-lateral movement, to a maximum of 100 km (Pindell
et al., 1998; Cerón-Abril, 2008). Assuming that the Oca fault accom­
The Guajira Peninsula is a tectonic block divided into two geological modated part of the Caribbean Plate motion (Pindell et al., 1988; Gómez
provinces, the La Alta Guajira Basin and the La Baja Guajira Basin, by the and thesis, 2001), Feo-Codecido (1972) proposed activity of the Oca
Cuiza fault, a large dextral strike-slip fault (Raasveldt, 1956; Renz, 1960; fault around the Eocene, beginning with mainly vertical activity and
MacDonald and Opdyke, 1972, Fig. 2). The peninsula is bounded to the changing to a dextral movement later during the late Eocene-Early
south by the Oca fault, a second right lateral strike-slip fault (e.g. Rod, Oligocene. On the other hand, Kellogg (1984) proposed the first move­
1956; Pindell, 1993; Audemard et al., 1994; Maceralli, 1995; Audemard, ment of the Oca fault during the Middle to Late Oligocene. Miocene
1996, Fig. 2). The origin of these structures is associated with the activity of the Oca fault has also been documented (Kellogg, 1984;
northeast to east right-lateral movement of the Caribbean Plate along Pindell and Kennan, 2009; Piraquive et al., 2017). Furthermore, Aude­
with the South American Plate (Pindell et al., 1998; Gomez and thesis mard (1996) considered that the Oca fault has been active during the
2001). The La Baja Guajira Basin is bounded by the western border of the Holocene as a high-angle reverse fault with a dominant dextral
Gulf of Venezuela eastward and by the South Caribbean Deformed Belt component. The geological map published by the Colombian Geological
(SCDB) westward. The Cuiza and Oca faults represent its northern and Service displays the trace of the Oca Fault cutting Neogene and Qua­
southern limits, respectively (Figs. 1 and 2). ternary sediments suggesting recent activity (Fig. 2).
The Oca fault marks the boundary between the Sierra Nevada de The Cuiza fault is considered a right-lateral strike-slip fault that also
Santa Marta (SNSM) and the La Baja Guajira Basin (Figs. 1 and 2). The accommodated the relative eastern movement of the Caribbean Plate
eastern extent of the fault covers ca. 650 km towards western Venezuela with respect to the South American Plate (Gómez and thesis, 2001). The
(Audemard, 1995). The westernmost termination of the Oca fault is western fault’s extent covers ca. 80 km of the offshore domain of the
traced against the northern extension of the Santa Marta-Bucaramanga Guajira Peninsula and the entire onshore domain where its strike dis­
fault along the offshore domain (Cerón-Abril, 2008, Fig. 1). Different plays a minor east-southeast orientation (Vence and Mann, 2020). The
authors have estimated the displacement of the Oca fault, going from 15 eastward limit of the Cuiza fault is proposed to be ca. 50 km into the Gulf

Fig. 2. Study area in detail with the main tectonic features. The base geological map is taken from Gómez et al. (2015). The Guajira Peninsula is divided into La Alta
Guajira and La Baja Guajira basins by the Cuiza Fault. Seismic and well data used for this study are also shown, which were kindly provided by Hocol S.A. The
location of the seismic sections shown in Chapter 4 is found here. The extension of the Tayrona Sub-basin is as proposed by Londoño et al. (2015). SCDB = South
Caribbean Deformed Belt; PD = Portete Depression; JR = Jarara Range; MR = Macuira Range; CHD = Chimare Depression; CD = Cocinetas Depression; CR = Cosinas
Range; SNSM = Sierra Nevada de Santa Marta.

3
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

of Venezuela (Vence and Mann, 2020). Estimations of lateral displace­ reported in the easternmost borehole Well-9 (Figs. 2 and 3) and there­
ment of the Cuiza fault vary from less than 1 km (Rollins, 1965) to 15 km fore excluded from our interpretation.
(Alberding, 1957; MacDonald, 1964; Gómez and thesis, 2001). Previous The basement in the La Baja Guajira Basin displays a complex
studies have reported the activity of the Cuiza fault starting during the assemblage of igneous, volcanic, and different-grade metamorphic rocks
Late Eocene until the Early Oligocene (Renz, 1960; Rollins, 1965; (Londoño et al., 2015; Mora-Bohórquez et al., 2017). In the boreholes
Alvarez, 1971; Gómez and thesis, 2001). It is proposed that reactivation considered for this study, the basement comprises quartzite and gneiss of
of the fault occurred during the Early Miocene, which displayed vertical continental affinity, amphibolite, and schist; and an igneous part that
offset in the offshore domain (Gómez and thesis, 2001). The activity of consists of diorite is described locally in the south-easternmost region
the Cuiza fault seemed to show minor intensity toward the Late (Well-8, Figs. 2 and 3). Radiometric ages reported for the basement
Miocene-Early Pliocene (Gómez and thesis, 2001; Vence and Mann, rocks in La Baja Guajira Basin are Lower and Upper Cretaceous (Mor­
2020), and younger movements are associated with vertical offset in the a-Bohórquez et al., 2017).
eastern extent of the Cuiza fault (Gómez and thesis, 2001). The Sediment deposited during the Cretaceous corresponds to dense
geological map displays the Cuiza Fault covered by Quaternary deposits limestone with fossiliferous levels, interbedded with calcareous shale
(Fig. 2). and marl of the Yuruma Formation (Valanginian-Berriasian, determined
The South Caribbean Deformed Belt (SCDB; Figs. 1–3) is an active, by Renz, 1956) (Vargas et al., 2012). Overlying this formation, the
sedimentary accretionary prism formed along the Caribbean–South Cogollo Group (Aptian-Cenomanian, determined by Renz, 1956) is
American margin due to the oblique subduction of the Caribbean Plate deposited, and it consists of thin silty calcareous and fossiliferous shale
underneath northern South America (Van der Hilst and Mann, 1994; layers interbedded with silty black limestone (Vargas et al., 2012).
Bernal-Olaya et al., 2015). It is characterized as to be a very thick wedge Upper Cretaceous sediment mainly comprises marine black shale with
of upper Miocene–Holocene piggyback deposits (Sanchez et al., 2019). interbedded fossiliferous limestone (Reilly, 1958) of the La Luna For­
The Chuchupa Platform is located south of the Cuiza fault (Fig. 3). It mation (Coniacian-Santonian, determined by Rollins, 1965). It should
is proposed to be part of the buried Great Arc of the Caribbean (Vence be noted that the La Luna Formation represents the main source rock for
and Mann, 2020). It consists of a dome-shaped basement high that ex­ hydrocarbon generation in northern South America (Lugo and Mann,
tends westward to the present shelf edge, and eastward it reaches the 1995), and in the eastern part of the La Baja Guajira Basin, this forma­
Ballena field (Gómez and thesis, 2001). The platform was likely tilting to tion contains traces of hydrocarbons (Ramirez et al., 2015). Further­
the east during the Oligocene, and to the west and south after the Early more, stratigraphic traps in the study area are of Cretaceous age and
Miocene (Gómez and thesis, 2001). Subtle movements reported during consist of onlapping reservoir units against basement highs (Cediel and
the Late Miocene to the Pliocene are restricted to the offshore domain, Aguilera, 2011, Fig. 4).
close to the shoreline, and related to an NW-SE-oriented young fault The Neogene rocks of the La Baja Guajira Basin include sandstone
system (Gómez and thesis, 2001). with intercalations of shale and biomicritic limestone (Cediel and
The Dorado High is a structural high located north of the Cuiza fault Aguilera, 2011) of the Uitpa Formation (Early Miocene, dated by Roll­
(Fig. 3), which is interpreted to result from Early Cenozoic faulting ins, 1965). The overlying formation corresponds mainly to fossiliferous
(Gómez and thesis, 2001). Fault activity during the Late Miocene had sandy limestone interbedded with calcareous sandstone (Vargas et al.,
been reported to the southwest (Hosie and thesis, 1994), and the west 2012) of the Jimol Formation (Lower to Middle Miocene, dated by
and south of the Dorado High (Duque-Caro and Reyes, 1999a,b). The Rollins, 1965). The Upper Miocene to Pliocene sediment (dated by
Dorado High was already a positive area during the Middle Miocene, Rollins, 1965) in the study area includes deepwater shale deposits
based on the termination (pinch out) of Middle Miocene seismic re­ intercalated with sandstone layers of the Castilletes Formation (Ramirez
flectors around it (Gómez and thesis, 2001). et al., 2015). Potential organic-rich source rocks in the study area
The Tayrona Sub-basin is an elongated depression located along the correspond to the shaly portion of the Castilletes Formation, dominated
offshore domain of the La Baja Guajira Basin and adjacent to the SCDB by kerogen type III, and it locally has reservoir potential in siliciclastic
(Figs. 2 and 3). Its sedimentary record comprises sequences deposited turbidite deposits (Ramirez et al., 2015, Fig. 4). The limestone sequence
from the Early Miocene to recent (Escalante, 2006). The Tayrona of the Jimol Formation is also determined as source rock in the La Baja
sub-basin is considered a forearc basin formed during the ongoing Guajira Basin (Barrero et al., 2007; Cediel and Aguilera, 2011). More­
interaction of the Caribbean Plate beneath northern Colombia that over, the sandy facies of the Uitpa Formation constitutes the lower
generated shortening deformation along the SCDB (Vence and Mann, reservoir in some of the wells in the Chuchupa field (Ramirez and thesis,
2020). Escalante (2006) interpreted two main periods of tectonic sub­ 2007), and the calcareous sandstone sequence of the Jimol Formation is
sidence and high sedimentation within the Tayrona Sub-basin, Early to considered the main gas reservoir in the same field (Ramirez et al.,
Late Miocene, and Late Miocene to recent, respectively. Interaction 2015). The Uitpa and Jimol formations act as seal rocks for Oligocene
between the Caribbean and South American Plates is responsible for and older reservoir rocks, while the shaly sequence of the Castilletes
periods of non-deposition or erosion from the Early Miocene to recent Formation is defined as the seal rock for Miocene reservoirs (Barrero
(Escalante, 2006). et al., 2007, Fig. 4). Furthermore, the structural traps present in the area
are Neogene in age and associated with thrust faults and near-vertical
2.2. Stratigraphic framework faults with dominant lateral displacement (Barrero et al., 2007; Ram­
irez et al., 2015, Fig. 4). In addition, stratigraphic traps associated with
The stratigraphy of the La Baja Guajira Basin mainly consists of Early Miocene sandstone reservoirs are proposed as the most effective
Miocene to Quaternary rocks (determined based on the biostratigraphic play on the Chuchupa Platform (Vence and Mann, 2020).
reports of the boreholes considered in this study, see Chapter 3) Finally, during the Pleistocene, the Gallinas Formation was depos­
deposited in marine and continental environments (Case et al., 1984). ited (Rollins, 1965), which consists of sandstone and claystone (Vargas
Triassic and Jurassic rocks are only reported in the northeastern part of et al., 2012; Ramirez et al., 2015). Although young and shallow, these
the study area, in the Cosinas Range, as part of the Guajira Arch (an rocks are also considered to act as potential reservoir rocks in the area
ENE-WSW trending high-grade basement intruded by igneous and (Ramirez et al., 2015, Fig. 4).
meta-igneous groups and covered by sedimentary rocks; i.e. Alvarez, Previous studies have interpreted the origin of the hydrocarbons in
1967; Zuluaga et al., 2015), and south of the Oca Fault within the Sierra the Guajira Basin mainly as microbial coming from a single petroleum
Nevada de Santa Marta (SNSM - Fig. 2; Gómez and thesis, 2001). system (Rice and Claypool, 1981; Katz and Williams, 2003; Escalante,
However, these units are not reported in the studied boreholes and are 2006) due to the molecular and isotopic composition of the gases and
thus not included in this paper. Cretaceous formations are only locally the rather low geothermal gradients measured in the area (Rangel et al.,

4
L. Castro-Vera et al.
5

Journal of South American Earth Sciences 121 (2023) 104109


Fig. 3. Regional-scale cross-sections showing the first-order structures and the distribution of the main depocenter in the study area. A) Section with a general NW-SE orientation. Chuchupa Platform, Tayrona Sub-basin
and what we interpreted as part of the South Caribbean Deformed Belt (SCDB) are shown in this figure. The location of the section corresponds to the red line in the map. B) Section with a general N–S orientation. The
Chuchupa Platform and the Dorado High are displayed here. The location of the section corresponds to the green line in the map.
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Fig. 4. Representative stratigraphic chart of the La Baja Guajira Basin. Geological formations, lithological descriptions, and petroleum system elements (PSE) are
modified after Cediel and Aguilera (2011). Triassic and Jurassic rocks are not included in this chart since they are not penetrated by the considered boreholes, and are
locally present in the Cosinas Range (further information can be found in i.e. Alvarez, 1967; Zuluaga et al., 2015). Correlation with seismic reflection data of the
study area and the interpreted seismic units (Acoustic basement-U7) are shown. Tops of U1, U2, and U3 are unconformities-check. U6 and U7 are not visible within
the seismic section since they are covering only the offshore area of La Baja Guajira Basin. See Fig. 2 for the location of the seismic line and Well-5.

2003). Other authors suggest the presence of two petroleum systems America break-up, which allowed the creation of a rift system,
(Cediel and Aguilera, 2011; García et al., 2008) and even four petroleum composed of several half-graben structures (Burke et al., 1984; Pindell
systems (Ramirez et al., 2015), considering both microbial and ther­ and Barrett, 1990; Lugo and Mann, 1995; Pindell and Kennan, 2009).
mogenic hydrocarbon generation. After rifting, a large passive margin was established along which
According to Ramirez et al. (2015), source rocks of Middle to Upper kilometer-thick marine sediments accumulated particularly in the
Miocene age are immature since they have not reached a sufficient western Lake Maracaibo (e.g. Mann et al., 2006). In the latest Creta­
temperature to produce thermogenic hydrocarbons. The organic matter ceous, the Caribbean plate moved in an easterly direction with respect to
of these rocks has been exposed to microbial activity since the Late the North and South American plates (Burke, 1988; Pindell and Barrett,
Miocene resulting in the generation of microbial methane accumulated 1990; Mann, 1999). An oblique collision at the Great Arc of the Carib­
in commercial volumes in the study area (Ramirez et al., 2015). The bean (GAC, defined by Burke, 1988) resulted in the formation of a
mixed isotopic signature of gas found in the Chuchupa field proves the foreland basin filled by an elongate sedimentary system extending from
existence of both microbial and thermogenic methane-rich gas there southern Colombia to the western Guajira Peninsula and Maracaibo
(Ramirez et al., 2015). Basin (MacDonald, 1964; Villamil, 1999).
During the Paleogene, oblique convergence between the Caribbean
2.3. Tectonic evolution Plate and South American passive margin produced convergent defor­
mation of the Guajira Peninsula (MacDonald, 1964). In the Oligocene,
The Guajira region and its offshore margins record a complex the surrounding areas (i.e., Serranía de Perijá) uplifted probably because
geological history between the Caribbean Plate and South American of shallow subduction of the Caribbean Plate on the South Caribbean
Plate involving plate collision (Paleogene), transform activity associated Deformed Belt (Taboada et al., 2000; Duerto et al., 2006; Gorney et al.,
with east-west-trending movement (Early-Middle Miocene; Audemard, 2007). Strike-slip faulting and an intense Andean uplift followed the
2009), and transtensional right-lateral strike-slip fault systems (post-­ Paleogene period during the Neogene (Mann et al., 2006). During the
Middle Miocene; Gorney et al., 2007). The Caribbean evolution started Middle Miocene, an increase in subsidence occurred in some Caribbean
during the Late Jurassic–Early Cretaceous by the North and South basins (i.e., western Falcon Basin; Escalona and Mann, 2004). In the

6
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Pleistocene, the Caribbean-South America boundary became similar to unit represent unconformable contacts and were associated with tec­
the present-day configuration. tonic events. Lithological descriptions and stratigraphic ages from the
studied boreholes were also integrated in order to assign approximate
3. Methods and data ages to the previously selected marker horizons (Table 1, Fig. 4). Faults
were interpreted based on detecting lateral reflection truncations,
3.1. Geological and geophysical data changes in the thickness of the seismic units, and disrupted and offset
reflectors. Folds were identified in areas of considerable reflection
This work builds on the integration of a basin-scale 2D seismic bending.
reflection dataset with borehole data. The analyzed data sets were Finally, the seismic-reflection data in time and all interpretations
provided by Hocol S.A and include (Fig. 2): I) 931 km of 2D seismic were converted into depth domain using interval velocities derived from
reflection data that consists of twenty-four (24) lines that correspond to borehole information (check shot data) and synthetic seismograms. It
nine (9) seismic programs acquired from 1969 to 1999. II) Lithological was necessary to plot the vertical travel time from the different datum
descriptions, biostratigraphic data, sample analysis, and velocity sur­ (ms) against the receiver vertical depth from the system reference datum
veys of nine (9) boreholes located offshore and onshore. III) Composite (ft.) of the boreholes that included this information (Well-3, Well-5,
logs. IV) Electrical logs (Spontaneous Potential, Gamma Ray, Caliper, Well-6, Well-8, and Well-9; Fig. 2). The plots allowed us to evaluate
Sonic, Density, and Resistivity). V) Petrographic and geochemical data the trend followed by the different wells and decide whether to apply
such as vitrinite reflectance (Ro%), total organic matter (TOC), Tmax, constant velocities along the study area or divide it into different ve­
S1, S2, and S3 peaks, hydrogen index (HI), and oxygen index (OI). locity areas. Constant velocities were applied and check shot data from
Additional literature research has been done comprising public Well-5 (see Fig. 2 for the location of the well) was taken as the base to
geological reports, theses, and scientific publications. calculate interval velocities (Table 1). A depth model was constructed
using the marker horizons previously interpreted as the surface intervals
(Table 1). Quality control was carried out by plotting respective seismic
3.2. Seismic interpretation
marker horizons against well tops.
Isopach maps were built in depth (m) for each seismic sequence,
The academic license of Petrel v2022 (Schlumberger©) was used to
already correlated to geologic time intervals, in order to provide a
develop the seismic interpretation and the time-depth conversion.
regional interpretation of the distribution of the depocenters and iden­
Twenty-four (24) seismic-reflection lines (2D) in two-way travel time
tify the main structural features within the La Baja Guajira Basin. The
(TWT) were selected to build a regional composite grid used in this study
results of the seismic interpretation provided initial inputs for basin
(Fig. 2). Subsequently, the respective wells’ correlation to the seismic
modeling, with geologic control from well logs.
data was carried out using available check shot information. For wells
that lacked this information, check shot data from the nearest borehole
was used and projected instead. 3.3. Basin modeling
All seismic-reflection data is shown and was analyzed in TWT, with
downward increasing impedance contrasts depicted as blue-colored Basin and petroleum system modeling (BPSM) is defined as the dy­
reflectivity peaks, under the SEG polarity (zero phase). First, well in­ namic modeling of geological processes affecting sedimentary basins
formation and synthetic seismograms were used to constrain the top over geological time intervals (Bruns et al., 2013). It includes deposition,
basement reflector’s location within the seismic-reflection data. The top pore pressure calculation, compaction, burial, and temperature his­
of the acoustic basement was linked to metamorphic and igneous tories, and the calculation of calibration parameters such as vitrinite
basement rocks using the boreholes that penetrated such rocks. reflectance (Ro%) or biomarkers based on kinetic data (Hantschel and
Thereafter, conventional seismic-reflection interpretation of signifi­ Kauerauf, 2009). Temperature histories are calibrated by comparing
cant horizons tied to borehole data and faults was carried out. The ho­ measured and calculated data, i.e., vitrinite reflectance, fission tracks,
rizons represent the boundaries between the main tectonostratigraphic and corrected bottom-hole temperature (Senglaub et al., 2005). Results
surfaces (Table 1). The criteria used to define these boundaries was of numerical 1D-basin models in this work were used to provide accurate
based on reflection terminations and major continuity surfaces with estimates of the burial and thermal history linked to the main tectonic
strong impedance contrast. The seismic reflection terminations used to events previously interpreted from the seismic observations.
identify the lower boundaries were onlap and downlap, while truncation In this study, 1D numerical simulation was performed on four
was utilized to delineate the upper boundary. The tops of each seismic selected boreholes located on- and offshore the La Baja Guajira Basin
(Well-,1, Well-3, Well-4, and Well-6; Fig. 2) using an academic license of
PetroMod v.2022 (Schlumberger©). The input data for the model con­
Table 1
struction included geological ages for each stratigraphic unit, depth,
Input data for the velocity model construction.
thickness, duration, and thickness of erosive events, and lithology of the
Age of the horizon Horizon Interval Interval
units (Table 2). The chronostratigraphic subdivision and absolute ages
Depth name velocities (ft./ velocities (m/
s) s)
were determined based on the chronostratigraphic time scale of the
International Commission on Stratigraphy (2013). Thicknesses and li­
Late Pleistocene Top U7 4572 1394
thologies were assigned based on the borehole reports provided by
Late Pliocene to Top U6 5500 1676
Early Pleistocene Hocol S.A. For determining the magnitude and timing of deposition and
Late Miocene to Top U5 6700 2042 erosive events, results from the seismic interpretation accomplished in
Pliocene this study were considered, as well as previous thermal models devel­
Late Miocene Top U4 7700 2347
oped for the study area (i.e., Escalante, 2006; García et al., 2008; Cediel
Middle Miocene Top U3 8200 2500
Early Miocene Top U2 8500 2591
and Aguilera, 2011; Ramirez et al., 2015; Vence and Mann, 2020).
Early Miocene Top U1 9200 2804 Additional information from the literature regarding the
Pre-Miocene Acoustic 9600 2926 tectono-sedimentary evolution was also integrated.
basement Once the conceptual model was validated, and the model’s input was
Pre-Miocene Base of the 14,000 4267
created, the thermal boundary conditions were defined, represented by
model
the basal heat flow, the sediment/water interface temperature (SWIT),
and the paleo-water-depth (PWD). Based on the information about
depositional environments given by the biostratigraphy reports of the

7
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Table 2 the acoustic basement to unit U7 (Figs. 4–6). The succession above the
Input data for 1D modeling of burial and temperature history of Well-6. acoustic basement is thickening westward reaching its maximum
Age Stratigraphic Depth Thickness Lithology thickness of ±4.0 s TWT (ca. 3000 m) within the Tayrona Sub-basin
(Ma) Tops/Well Picks [m] [m] (Fig. 5). Northward, the succession thins to a minimum thickness of
0.10 Upper Pliocene 0 305 Sandstone (40%), ±0.8 s TWT (ca. 500 m) where the basement forms prominent paleo-
to younger limestone (50%), coal highs (Dorado High and Chuchupa Platform; Fig. 5). Seismic units U1
(10%) to U7 are interpreted to be deposited between the Early Miocene and
3.60 Lower Pliocene 305 671 Siltstone (40%), Pleistocene based on biostratigraphic data from the boreholes consid­
sandstone (30%),
mudstone (25%), coal
ered in this study (Figs. 2 and 4).
(5%) Acoustic basement. The top of the acoustic basement corresponds to
5.30 Upper Miocene 975 884 Clay-rich silic. mudstone a high-amplitude reflector that deepens westwards from ca. 0.8 s TWT in
(40%), siltstone (40%), the proximal area (today’s onshore La Baja Guajira Basin) to up to 7 s
sandstone (20%)
TWT toward the distal margin (along the Tayrona Sub-basin; Figs. 5 and
11.62 Upper Middle 1859 792 Shale (55%), siltstone
Miocene (35%), sandstone (10%) 6). A general basinward thinning of the basement can be inferred by this
15.97 Lower Miocene 2652 292 Limestone (80%), shale rapid and significant deepening. Boreholes document that the top of the
(10%), siltstone (10%), acoustic basement consists of igneous and metamorphic rocks underly­
23.00 Basement 2944 10 Quartzite ing sedimentary rocks (Table 3). The acoustic basement is affected by
normal faults trending NW-SE, a relict to extensional deformation,
boreholes, paleo-water depths were determined. In general, for Lower which resulted in the formation of horst, graben, and half-graben
Miocene deposits, inner to middle shelf conditions were assumed. Outer- structures (Figs. 5 and 6). Prominent basement lows correspond to the
shelf to bathyal environments were assigned for Middle to Upper main depocenters of the study area (Figs. 5 and 6). Basement highs are
Miocene deposits. For defining the Pliocene to Pleistocene PWD, water associated with major normal faults that in some cases show an inver­
depths at the present-day configuration of the basin were taken into sion. The most noticeable basement highs are the Dorado High and
account. The SWIT for all events were then calculated based on paleo- Chuchupa Platform located in the proximal offshore, north and south of
latitude and respective paleo-water depths of the study area (Wygrala, the Cuiza Fault system, respectively (Fig. 5).
1989). Seismic unit U1. The base of unit U1 corresponds to the top of the
The heat flow history was set as simply as possible; this means a acoustic basement. The top of U1 is characterized by a dominantly high-
similar heat flow trend within the study area has been assigned amplitude reflection that represents an unconformity (Fig. 4). U1 is
following the typical heat flows associated with different sedimentary thickening (up to 300 m thick; Fig. 7) to the east of the Chuchupa
basin types as defined by Allen and Allen (2013). The heat flow values Platform. It shows in places downlaps of overlying reflections onto U1’s
for the period of maximum temperatures (i.e., Middle to Late Miocene) top (Fig. 8). U1 contains wedge-shaped packages of sub-parallel and
were calibrated using vitrinite reflectance data; therefore, they are a continuous to semi-continuous medium amplitude reflectors that onlap
result of the modeling approach. The heat flow trends were derived onto prominent basement highs (i.e. Dorado High; Figs. 5 and 9). U1 is
based on different research such as Apatite Fission Track and Zircon deepening westward from ca. 0.8 s TWT along today’s onshore La Baja
Fission Track (García et al., 2008), vitrinite reflectance (García et al., Guajira Basin to up to 3 s TWT towards the offshore area (Figs. 5, 6 and
2008; Instituto Colombiano del Petróleo, 2000), and downhole tem­ 8). U1 onlaps onto the Dorado High to the north and downlaps westward
perature data (Escalante and thesis, 2005; García et al., 2008; internal against the Chuchupa Platform (Figs. 5 and 6). Within the onshore area,
borehole reports of Hocol S.A). U1 exhibits humps with hummocky reflection patterns, features that are
The thermal and burial histories of the selected wells were calibrated not linked to structures of the underlying basement, and that can be
by comparing measured and calculated vitrinite reflectance (Ro%) and interpreted as prograding flanks of carbonate platforms (Fig. 8).
corrected bottom-hole temperature data. The EASY %Ro algorithms Microscopic analyzes (Perez et al., 1988) indicate that the calcareous
(Sweeney and Burnham, 1990; Burnham, 2016, 2019) were selected for sequence drilled in Well-5 (correlated to the depth at which this hump
the calculation of vitrinite reflectance. The EASY %Ro models incorpo­ structure is observed in the seismic) corresponds to a platform carbonate
rate four reactions (generation of water, carbon dioxide, higher hydro­ deposit in a low-energy environment.
carbons, and methane) in one activation energy distribution that allows U1 has an earliest Early Miocene age based on biostratigraphic data
the calculation of vitrinite reflectance values between 0.3 and 4.5 Ro% from boreholes that reached this unit (Well-5, Well-8, Well-4, Well-6;
(Rodon and Littke, 2005). Regarding the borehole temperatures, the internal geological well reports). U1 can only be observed along to­
values were corrected using 10, 15, and 20% of correction. Measured day’s onshore and proximal offshore areas, and it is thus considered a
vitrinite reflectance and temperature data are the basis for the calibra­ restricted event when compared to the overlying units (Fig. 7).
tion of our 1D simulations of burial and thermal history. Seismic unit U2. Unit U2 is bounded to the base by the top of unit U1
A representative input dataset of the Well-6 (see location in Fig. 2) is and to the top by a semi-continuous low amplitude reflector. The top of
presented in Table 2, containing data of layer thickness, lithology, age, this tectonosequence is also an unconformity (Fig. 4) and is deepening
stratigraphic borehole picks, and top depth for each unit. westward from ca. 0.2 s TWT in the onshore area to ca. 6 s TWT in the
In addition, since sedimentation is considered one of the primary distal offshore area (Figs. 5 and 6). U2 is characterized by semi-to
controlling factors for hydrocarbon accumulation and preservation discontinuous reflections of medium frequency and low amplitude,
(Katz, 2001), sedimentation rates were estimated based on the chrono­ high to medium amplitude reflectors are locally observed within the
logical and thickness information from different geological formations western depocenter (Tayrona Sub-basin; Figs. 5 and 10). U2 thickens up
in the basin. to 800 m in the Tayrona Sub-basin (Figs. 5, 7 and 10). In the southeast,
significant thickening of this unit is controlled by normal faults (possible
4. Results growth strata; Figs. 6 and 11). Anticlines locally deforming this sequence
are seen on the eastern side of the Tayrona Sub-basin and on eastern
4.1. Seismic interpretation Dorado High (Figs. 5 and 10). On the western flank of the Tayrona Sub-
basin, unit U2 is displaying folding deformation (Figs. 5 and 10).
4.1.1. Seismic units and age correlation The age of U2 is interpreted as latest Early Miocene based on the
The studied succession was subdivided into eight (8) seismic units, biostratigraphic data reported from samples analyzed (Well-5, Well-8,
Well-9, Well-7, Well-1, and Well-3; internal geological reports). U2 is

8
L. Castro-Vera et al.
9

Journal of South American Earth Sciences 121 (2023) 104109


Fig. 5. 2D seismic section with a general orientation W-E. A) Uninterpreted section. Well-2 and Well-3 reach the basement. B) Interpreted section that displays the acoustic basement and the seismic units determined in
this study. Fault Group 1 is observed affecting the basement and the syn-extension sequences U1 and U2. Fault Group 2 corresponds to the inverted rooted-basement normal faults. Faults of group 2 are linked to the
inversion anticlines, shortcut faults, uplift of the Dorado High, and folding of the seismic units. The younger seismic units U5 and U6 display the characteristic westward prograding clinoforms that have already entered
the proximal offshore area. C) Zoom into the western flank of the Tayrona Sub-basin where we detected BSRs possibly associated with methane hydrates.
L. Castro-Vera et al.
10

Journal of South American Earth Sciences 121 (2023) 104109


Fig. 6. 2D seismic section with a general orientation W-E. A) Uninterpreted section. Well-5, Well-8 and Well-9 reach the basement. B) Interpreted section that displays the acoustic basement and the seismic units
determined in this study. Structural and stratigraphic features detected in the area are shown. Growth strata of U2 controlled by normal faults (Fault Group 1) are observed towards the southeast. Seismic units U3, U4,
U5, and U6 display the characteristic westward prograding clinoforms within the onshore and proximal offshore areas. Steep faults (Fault Group 3) interpreted to have a strike-slip component, are affecting the Neogene
units along the onshore area.
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Table 3 to 570 m thick; Fig. 7) in the southcentral area, this thickness is related
Lithological description of the boreholes drilling crystalline rocks associated to the accumulation of clinoforms (Figs. 8 and 12). As in the underlying
with the basement in the study area. See Fig. 2 for the location of the boreholes. U3, these clinoforms are prograding westward (Figs. 6 and 8). It is
Well Top Description Reference important to mention that along the areas where U3 is thickening (i.e.,
name depth depocenter northwestward), unit U4 is thinning (Fig. 7). U4 thins locally
[m] at the top of the Dorado basement high (Fig. 9), and in the same northern
Well-6 2900 Metamorphic. Dark green olive, with Duque-Caro and direction, U4 is truncated by an overlying event (U7; Fig. 5). The re­
abundant serpentine and amphibole, Reyes (2000) flections of U4 are continuous to semi-continuous with a medium fre­
very weathered red film.
quency and predominantly low amplitude, the amplitude becomes
Well-4 2080 Metamorphic. Amphibolite. Fine- Western Atlas
grained, dense, greenish grey, slight International, 1990 higher at the base of the unit and at the topset of the clinoform struc­
schistosity. Dominant minerals are tures. The seismic pattern that characterizes U4 does not show much
hornblende, plagioclase feldspar, significant faulting control and its thickness remains rather constant.
quartz and iron. Schist (?). Along the western Tayrona Sub-basin, U4 is folded (Figs. 5 and 10).
Well-3 1660 Metamorphic. Grey schist. Bluish- Duque-Caro and
grey quartzose phyllite. Reyes (1999)
U4 can be interpreted as a Late Miocene event based on the
Well-9 2200 Dark green, greasy, highly siliceous Reilly (1958) biostratigraphic data documented by the wells that reach this unit (all
igneous to metamorphic rock, the studied wells; internal geological reports). U4 is a widespread
tentatively identified as serpentinite. sequence (Fig. 7).
Well-2 2940 Metamorphic. Quartzite. Angular Texas Petroleum
Seismic unit U5. The base of U5 corresponds to the top of unit U4
fragments of dark grey, milky white, Company, 1989
and dark brown crystalline rocks. and its top represents a continuous reflectivity peak with a predomi­
Well-8 2440 Igneous. Dark and green phaneritic Lawrence (1958) nantly low amplitude that becomes medium along the offshore area. The
igneous with 60–90% red and green top of this sequence is deepening westward from ca. 0.2 s TWT within
claystone, which is assumed to be the onshore area to ca. 4.6 s TWT in the offshore distal area (Figs. 5 and
caving. The one-foot recovery was
10). There are downlapping terminations onto U5’s top (Figs. 8, 11 and
dark grey to green igneous (diorite).
Well-5 2860 Metamorphic. Clear, translucent Ecopetrol (1988) 12). U5 is thickening in the proximal offshore area (up to ca. 480 m
quartzite, composed of medium to thick; Fig. 7) where it displays clinoforms (Figs. 8, 11 and 12). The cli­
coarse-grained quartz, locally noforms show the continuation of the westward progradation observed
conglomeratic in appearance,
since seismic event U3, and downbuilding of clinoforms is observed at
angular to sub-rounded, with biotite
inclusions that present some the base of U5 (Fig. 12) indicating a loss in accommodation. Like U4, the
orientation giving a slight shale seismic event U5 shows thinning within the structural depocenter
appearance, also present traces of (Tayrona Sub-basin; Figs. 5 and 7). Northwards, the younger seismic
muscovite. event U7 truncates U5 (Fig. 5). The seismic character of unit U5 is very
similar to that of U4.
a widespread sequence with sediments deposited in distal offshore areas U5 is interpreted as Late Miocene to Pliocene based on biostrati­
(Figs. 5–7). graphic data reported by wells drilling this unit (all studied wells; in­
Seismic unit U3. The base of this unit corresponds to the top of unit ternal geological reports). In the northwest, submarine canyons are
U2. The top of U3 is a continuous reflectivity peak with a medium observed within unit U5 suggesting erosive events affecting the basin at
amplitude eastward and higher amplitudes westward (more distal area). least along the distal offshore area (Fig. 5). U5 shows less folding along
The top of this sequence is deepening westward from ca. 0.3 s TWT the western flank of the Tayrona Sub-basin in comparison to U2, U3, and
within the onshore and Chuchupa Platform areas to ca. 5 s TWT in the U4 (Figs. 5 and 10).
distal offshore area (Figs. 5, 6 and 10). U3’s top is an unconformity Seismic unit U6. U6 is bounded to the base by the top of U5. It is
marked by a lateral truncation of underlying reflectors (Figs. 4, 5, 11 and bounded at the top by a semi-continuous reflection peak with a medium
12). There are downlaps observed on U3’s top (Figs. 8, 11 and 12). U3 to low amplitude. The top of U6 is deepening westward from ca. 0.1 s
consists of parallel to subparallel and semi-continuous reflections with TWT in the proximal offshore area to ca. 3.6 s TWT in the distal offshore
high amplitudes and high frequencies that become lower towards its top. area (Figs. 5 and 6). The top boundary is an unconformity in the areas
U3 displays the highest thickness of all the interpreted seismic units (up close to the Cuiza Fault (Fig. 5). U6 thickens westward and reaches its
to 1400 m thick; Fig. 7). It is thickening away from the Chuchupa maximum thickness at the Chuchupa Platform (up to 1100 m thick;
Platform, within the depocenters formed by structural basement lows, Fig. 7). In this area, westward trending clinoforms are well-documented
which are the same as for U2 (Figs. 5 and 7). U3 thins towards the central and downbuilding of clinoforms is observed (Figs. 5, 11 and 12). A high
study area at the Chuchupa Platform (Figs. 5, 11 and 12). In the thickness of U6 is also observed along the depocenter located in the
southeast (onshore area), U3 shows clinoforms prograding westward offshore distal area (Tayrona Sub-basin; Figs. 5, 7 and 10). The seismic
(Figs. 6 and 8). In the proximal and distal offshore areas, the clinoforms pattern that characterizes U6 is very similar to that of U4 and shows a
are not evident (Figs. 5, 6, 9 and 10). In the north, sequence U3 onlaps significant sediment input with massive clinoform accumulation
onto the Dorado High (Fig. 9), and its top is truncated by an overlying (Figs. 5, 11 and 12).
event (U7; Fig. 5). Along the western flank of the Tayrona Sub-basin, this The seismic unit U6 is interpreted as Late Pliocene-Early Pleistocene
unit is folded (Figs. 5 and 10). based on the biostratigraphic data documented in the offshore wells
U3 is interpreted as a Middle Miocene event based on biostrati­ (Well-2 and Well-3; internal geological reports). U6 has been only
graphic data from wells penetrating this unit (all studied wells; internal recognized in today’s offshore areas. In the current onshore La Baja
geological reports). The deposition of unit U3 does not show a direct Guajira Basin, this unit is at or near the present-day surface (see
fault-controlled pattern. As unit U2, seismic unit U3 is a widespread geological map in Fig. 2). Because of its shallow position, it is not imaged
sequence (Fig. 7). on the seismic-reflection data.
Seismic unit U4. The boundaries of U4 are the top of U3 at the base Seismic unit U7. The base of U7 is the top of unit U6. The top is the
and, at the top, a reflector with medium amplitudes that become higher present-day seafloor. Therefore, unit U7 extends only within today’s
locally in today’s proximal offshore area. The top of this seismic unit is offshore La Baja Guajira Basin (Figs. 5–7). As the underlying units, the
deepening westward from ca. 0.27 s TWT in the onshore area to ca. 5 s top of unit U7 is deepening westward from ca. 0.1 s TWT in the more
TWT in the distal offshore area (Figs. 5 and 6). U4 thickens the most (up proximal area to ca. 3 s TWT in the distal area (Fig. 5). U7 displays a
thickening trend like the underlying unit U6, with a maximum thickness

11
L. Castro-Vera et al.
12

Journal of South American Earth Sciences 121 (2023) 104109


Fig. 7. Isochore maps in meters of the interpreted seismic units. The maps are in the same scale and color code. The maps show the thickening and thinning trends of the seismic units as well as the main depocenters in
the study area. The onshore (OA), proximal offshore (POA), and distal offshore areas (DOA) are presented in the maps. The limit between the OA and POA is the continent line, and the POA and DOA are separated by the
grey dashed line (that represents the paleo-shelf break) located west of the Chuchupa Platform.
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Fig. 8. A) Uninterpreted seismic line with W-E orientation, located along the onshore area in the basin (see the location of the seismic line in Fig. 2). B) Interpreted
seismic section pointing out the termination relationships of the reflections and the seismic horizons interpreted as to have tectonic significance. Steep faults (Fault
Group 3) are shown cutting the sedimentary infill. Seismic units U3, U4, and U5 display the characteristic westward prograding clinoforms. C) Zoom into the domed-
shaped structure performed by unit U1 that is interpreted as a carbonate platform. Well-5 is also displayed with its gamma ray log that marks a clear change into
carbonate rocks when entering unit U1.

of ca. 560 m on top of the Chuchupa Platform in the proximal offshore (Early Miocene; Fig. 5); however, they also affect younger seismic units
area (Fig. 7). Erosional features (submarine canyons) are present within (Figs. 8, 10 and 11). The Dorado High and Tayrona Sub-basin are bound
seismic unit U7 in the northwest characterized by onlapping infill re­ by both non-inverted and inverted normal faults (fault groups 1 and 2;
flections (Fig. 5). U7 is interpreted as Late Pleistocene to recent based on Figs. 5 and 9).
the geological report of Well-1 (internal geological reports). Fault Group 3 - East-west oriented strike-slip faults: these fault
systems are the main bounding elements of the study area, the Cuiza and
4.1.2. Fault interpretation Oca faults (Fig. 2). On the interpreted 2D seismic data, strike-slip faults
The interpretation of seismic reflection data reveals at least three are difficult to observe; where imaged, these faults exhibit a small
main fault groups that affect the basement and the sedimentary infill. normal component that we correlated to steep planar faults (Fig. 8).
Among these groups, normal faults are the most common followed by Such faults were mainly observed near the Oca Fault in the south,
strike-slip faults. Some evidence of tectonic inversion was detected. affecting much of the sedimentary succession (Figs. 6 and 8).
Fault Group 1 - Basement normal faults: this group mainly con­
sists of planar northwest striking faults that have dips ranging from 4.2. Basin modeling
moderate to high-angle (Figs. 5 and 6). Basement normal faults are
distributed along today’s onshore and offshore La Baja Guajira Basin. Results obtained from the 1D basin modeling of four wells located in
These faults are interpreted as extensional structures that displace the the La Baja Guajira Basin are presented in the following section. The
acoustic basement, resulting in horsts, half-grabens, and graben struc­ input data originate from the Well-1, Well-3, Well-4, and Well-6 (see
tures (Figs. 5 and 6). These faults bound the depocenters of the study well locations in Fig. 2). The simulation output shows differences in
area, the Tayrona Sub-basin to the northwest (Fig. 10), a structural burial and thermal maturation history for the areas covered by the
depression located north between the Chuchupa Platform and the Do­ modeled wells (Fig. 13).
rado High (Fig. 9), and a minor structural low in the south-central part of
the study area (Fig. 7, units U2 and U3). 4.2.1. Burial history
Fault Group 2 – Inverted basement-rooted faults: this group The La Baja Guajira Basin exhibits two periods of rapid tectonic
consists of reactivated normal faults of Fault Group 1. Therefore, the subsidence and high sedimentation rate (Fig. 13 and Table 4). The first
geometry of the faults is similar to the description of Fault Group 1. period occurred during the Middle Miocene, and the second period is
Inversion can be indicated by local anticlines, structural uplift on the observed in the Early Pliocene. However, differences in the intensity and
hanging wall of normal faults, shortcut faults, and thickness variation duration of these events have been recognized within the study area.
over the limbs and core of the inverted anticlines (Figs. 5, 6 and 9). Therefore, the results presented below are divided between the north
Inverted basement-rooted normal faults mainly occur close to the Oca (Well-3 and Well-4) and south (Well-1 and Well-6) areas of the basin.
and Cuiza faults. These faults are most pronounced in units U1 and U2 In Well-1 and Well-6 (southern area), continuous and moderate

13
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Fig. 9. A) Uninterpreted seismic line with N–S orientation,


located along the offshore area in the Cuiza fault system at the
La Baja Guajira Basin (see the location of the seismic line in
Fig. 2). B) Interpreted seismic section displaying the main
basement highs in the study area, Chuchupa Platform and
Dorado High, and the terminations relationship of the seismic
units U1, U2, and U3 onto the Dorado High. Seismic unit U1
thickens the most within the structural low formed in between
the basement highs. U3 and U4 display thickness variation in
relation to the Dorado High, which we associate with the uplift
of the Dorado High, while no thickness variation is observed in
units U5, U6, and U7.

subsidence is observed from the Early Miocene (23.03 Ma; Fig. 13A and showing a possible natural cooling/tectonic quiescence of the basin with
B) while rapid tectonic subsidence occurs from the Middle Miocene a low angle slope until the present day at both locations (Fig. 13C and
(16.02 Ma) to the Late Miocene (5.34 Ma). Minor variations in the D).
subsidence rate are present during this period with an average of ca. 290 Calculated sedimentation rates range from ca. 9–900 m/my
m/my (Well-1) and ca. 420 m/my (Well-6). The second phase of faster (Table 4). They are highest in the Early Pliocene in all wells and also
tectonic subsidence occurred from 5.34 to 3.57 Ma (Pliocene) at a rate of quite high in the Middle and Late Miocene. Sedimentation rates are
ca. 460 m/my in Well-1 and ca. 750 m/my in Well-6. A period of tectonic highest in the southern offshore of La Baja Guajira Basin (Well-1;
quiescence is captured in Well-1 after the Pliocene (3.75 Ma) until the Table 4).
present day, while moderate subsidence (ca. 90 m/my) is captured by
Well-6. From the Early Miocene to the present day, a total of 4790 m of 4.2.2. Thermal and maturity history
sediments have been deposited as reported for Well-1, and 2940 m as Fig. 13 shows the thermal (temperature isolines in grey) and matu­
indicated by Well-6 data (Fig. 13A and B). rity (Sweeney&Burnham (1990) Easy%Ro model) histories of the
At the locations of Well-3 and Well-4 (northern area), subsidence modeled wells overlaid on the burial curve as well as the thermal and
started from the Early Miocene (23.03 Ma) at rates of 53 m/my and 17 maturity calibrations. The most suitable match between measured and
m/my respectively (Fig. 13C and D). The subsidence rate increased from calculated present-day temperatures was achieved with present-day
16.02 Ma until 11.66 Ma (Late Miocene) along the Well-4 location at basal heat flows in the range of 33–63 mW/m2. Calibration of the
rates of around ca. 820 m/my. Modeling at Well-3 indicates a change to models based on maturity data (vitrinite reflectance, Ro%) was
rapid tectonic subsidence for a short period during the Middle-Late accomplished by determining the heat flow values throughout time and
Miocene (from 11.62 to 11.4 Ma) when the subsidence rate reached considering the tectonic events affecting the study area.
ca. 980 m/my. Following the strong subsidence, a phase of decreasing During the Early Miocene (23–15 Ma), we assumed a constant basal
subsidence and sedimentation continued until the Pliocene (5.27 Ma). heat flow that ranges from 55 mW/m2 (Well-6), 65 mW/m2 (Well-1) to
Subsequently, the second episode of rapid tectonic subsidence occurred 70 mW/m2 (Well-3 and Well-4) based on the average heat flow sug­
at Well-4, while at Well-3 the subsidence rate decreased. This phase of gested for extensional basins by Allen and Allen (2013). From the Middle
rapid tectonic subsidence at Well-4 lasted until the upper Pliocene (2.65 to Late Miocene (15–10 Ma), basal heat flow values remained the same
Ma) at a rate of ca. 600 m/my. At 2.53 Ma, the subsidence decreased as during the Early Miocene for all the wells. We correlated these values

14
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Fig. 10. A) Uninterpreted seismic section with W-E


orientation, located along the distal offshore area of
the study area (see the location of the seismic line in
Fig. 2). B) Interpreted section. Termination relation­
ships between the reflections and the seismic units are
displayed. Evidence of inversion is observed towards
the west flank of the Tayrona Sub-basin and it is
associated with the reactivation of larger preexisting
basement-rooted faults (Fault Group 2). U2 thickens
the most within the Tayrona Sub-basin. The inter­
pretation of the sequences is not possible to be
continued to the west since the reflectors are not
clear.

Fig. 11. A) Uninterpreted seismic section with SW-


NE orientation, located along the proximal offshore
area of the basin (see the location of the seismic line
in Fig. 2). B) Interpreted seismic section showing the
terminations relationship of the reflections and the
seismic horizons interpreted as to have tectonic sig­
nificance. The internal geometry of the clinoforms
performed by the Neogene units U5 and U6 is high­
lighted. U6 reaches its maximum thickness along this
seismic line due to the topset accumulation.

with the typical heat flows of active strike-slip basins (Allen and Allen, we considered during this period are 33 mW/m2 (Well-6), 45 mW/m2
2013); they can be associated with recurrent reactivation of the Cuiza (Well-1 and Well 4), and 63 mW/m2 (Well-3). No erosion periods were
and Oca strike-slip faults. For all the wells, reduced basal heat flow assumed in order to reach the calibration of the maturity curves. It is
values were modeled from the Pliocene to present day. The values that important to mention that to achieve a suitable fit between measured

15
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Fig. 12. A) Uninterpreted seismic section with W-E orientation, located along the proximal offshore area in the La Baja Guajira Basin (see the location of the seismic
line in Fig. 2). B) Interpreted seismic section showing the terminations relationship of the reflections and the seismic horizons. Seismic units U4, U5, and U6 display
the characteristic westward prograding clinoforms, and the internal geometry of these clinoforms is highlighted. C) Detailed clinoform interpretation along the
seismic line. Note distinct downbuilding of clinoforms at the base of units U5 and U6. Downbuilding clinoforms are independent of sedimentation and always indicate
a loss in accommodation space (Catuneanu et al., 2009).

and calculated vitrinite reflectance data of Well-3 and Well-4, very high events, as well as observations derived from analyzing temperature and
heat flow values (>90 mW/m2) and/or a high amount of eroded sedi­ subsidence curves.
ments (>2 km) are required during the Pliocene; however, we consider
these conditions as geologically unrealistic for the study area.
5.1. Pre-extension sequence
The thermal modeling indicates that all the rocks in the study area
have been subjected to the maximum temperatures approximately for
The acoustic basement mainly consists of crystalline rocks and
the last 6 Ma (Fig. 13). The Middle Miocene source rocks (Jimol For­
metasediments (i.e., Gorney et al., 2007; Londoño et al., 2015; Vence
mation) have reached the highest temperatures (up to 91 ◦ C) in recent
and Mann, 2020) deposited prior to an extension phase and therefore
times along the south offshore La Baja Guajira Basin (Well-1), while the
affected by northwest-striking normal faults (Family Group 1). The
Upper Miocene source rocks (Castilletes Formation) have experienced
extension event that affected the basement and that occurred prior to the
up to 70 ◦ C in the same area (Well-1; Fig. 13A). Even though tempera­
deposition of the oldest seismic unit interpreted in this study (unit U1),
tures reached at present time have been just sufficient to enter the oil
can be correlated with an Oligocene transtension phase (Gorney et al.,
window, the source rocks have remained immature with Ro% values
2007; Escalante, 2006; Vence and Mann, 2020) and supported by
ranging from 0.37 to 0.55 (based on our model). Modeling at Well-1 and
right-lateral strike-slip faulting during this period (Escalona and Mann,
Well-6 shows that the present-day conditions (ongoing subsidence)
2011). This transtension resulted in the formation of basins (Escalona
would lead middle Miocene source rocks (Jimol Formation) to reach
and Mann, 2011; Vence and Mann, 2020), and probably the horst and
further maturation.
graben structures (i.e. Dorado High and Chuchupa Platform).
Radiometric ages of basement samples within the Guajira Peninsula
5. Geological interpretation and discussion
range from 47 to 128 Ma (K–Ar, 40Ar/39Ar, LA-ICP-MS; Vence, 2008),
ages that coincide with the ones compiled from the Great Arc of the
The seismic units, burial, and thermal models interpreted within the
Caribbean (GAC; Vence, 2008). Different authors (Pindell et al., 1998;
La Baja Guajira Basin record the Neogene and Quaternary structural and
Gomez, 2001; Montes et al., 2005; Cardona et al., 2007; Gorney et al.,
sedimentary evolution that contributes to the understanding of the
2007; Vence and Mann, 2020) support the presence of the GAC (an Early
Cenozoic development of the northern Colombian margin. Key Neogene
Cretaceous to recent volcanic arc), at least along the offshore Guajira
to Quaternary geological events are summarized in Fig. 14 and discussed
area, which can also be considered as a discontinuous block in the
below in chronological order. They include depositional and structural
Guajira peninsula as the result of syn- or post-suturing deformation

16
L. Castro-Vera et al.
17

Journal of South American Earth Sciences 121 (2023) 104109


Fig. 13. 1D plots indicating the thermal and burial history, as well as the calibration curves (Temperature and Vitrinite Reflectance) at the location of the modeled boreholes. Sweeney&Burnham (1990)Easy%Ro model
is used as an overlay on the burial plots to indicate the maturity. Grey lines on the burial plots correspond to isolines of temperature. The episodes of rapid tectonic subsidence are shown in figures. A) Well-1, B) Well-6,
C) Well-3 and D) Well-4.
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109
Fig. 13. (continued).
18
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Table 4 southwestern end of the Chuchupa Platform - Fig. 11), indicates the
Age information, the thickness of the intervals reported in the modeled wells, activity of northwest-striking normal faults (Fault Group 1). Gorney
and calculated linear sedimentation rates (LSR). et al. (2007) interpreted as well extension motion along
Well Interval Age top Age Thickness LSRa northwest-striking normal faults between the Leeward Antilles islands
name (my) bottom (m) (m/my) during the Early Miocene.
(my) Previous studies carried out to the south domain of the Oca Fault
Well-6 Upper Pliocene 0,1 3,6 305 87 (Palomino basin), stated extensional faulting with the onset of dextral
to younger strike-slip activity of the Oca fault in the Miocene as a consequence of
Well-6 Lower Pliocene 3,6 5,3 670 394
the interaction between South American and Caribbean crusts (Pira­
Well-6 Upper Miocene 5,3 11,62 884 140
Well-6 Upper Middle 11,62 15,97 792 182 quive et al., 2017).
Miocene
Well-6 Lower Miocene 15,97 23 292 42 5.3. Middle to Late Miocene: post-extension and localized inversion
Well-4 Pleistocene 0,01 2,58 25 9
Well-4 Pliocene 2,58 3,6 304 298
Well-4 Lower to Middle 3,6 5,3 735 432
The units U3 and U4 represent post-extension sequences (Fig. 14).
Pliocene Significant residual accommodation generated during the Early Miocene
Well-4 Upper Miocene 5,3 11,62 302 48 extension started being filled from the Middle Miocene on. This period is
Well-4 Middle Miocene 11,62 15,97 527 121 characterized by oceanwards prograding clinoforms (Vence and Mann,
Well-4 Lower Miocene 15,97 23 180 26
2020, Figs. 6 and 8).
Well-3 Upper Pliocene 2,58 5,3 262 96
Well-3 Upper Miocene 5,3 11,63 872 138 The considerable thickness of unit U3 (Fig. 7) marks an increment of
Well-3 Upper Middle 11,63 13,82 338 154 sediment input in the Middle Miocene, which was probably sourced by
Miocene the exhumation of adjacent mountain belts such as the Santa Marta
Well-3 Middle Miocene 13,82 15,97 43 20
massif and the Sierra de Perijá (Hosie and thesis, 1994; Gómez et al.,
Well-1 Lower Pliocene 3,6 5,3 1533 902
Well-1 Upper Middle 5,3 11,63 1527 241
2005; Bayona et al., 2007; Villagómez et al., 2011; Mora-Bohórquez
Miocene et al., 2017; Piraquive et al., 2017; Patiño et al., 2019). This exhumation
Well-1 Middle Miocene 11,63 15,97 1158 267 is considered a consequence of the convergence between the Caribbean
Well-1 Lower Miocene 15,97 23 275 39 and South American plates (Mora-Bohórquez et al., 2017; Piraquive
a
Linear Sedimentation Rate calculated as LSR = Thickness/(Age top - Age et al., 2017; Bermúdez et al., 2019), or a consequence of compression
bottom). originated in the Pacific margin of South America (Villagómez et al.,
2011; Patiño et al., 2019). The increment of sediment input resulted in
(Blanco et al., 2015). higher sedimentation rates and a significant rise in subsidence from the
The top of the basement abruptly deepens toward the distal margin Middle Miocene in the study area (Fig. 13). Maximum subsidence in the
and reaches larger depths (ca. 12 km depth) along today’s northwestern vicinity of the La Baja Guajira Basin has also been reported during the
offshore area (Tayrona Sub-basin), inferred in this study as a basinward Middle Miocene (i.e., Maracaibo Basin; Mann et al., 2006) being
thinning of the basement. This rapid decrease in crustal thickness might correlated with the beginning of the uplift of the Merida Andes east of
feature modern-day transform margin (continent-ocean transition; the Maracaibo Lake (Castillo and Mann, 2006). The continuous sedi­
Cerón-Abril, 2008; Londoño et al., 2015). mentation led rocks of the southern areas (models of Well-1 and Well-6)
to reach even higher burial depths during the Late Miocene (Fig. 14).
Yet there is evidence for at least two episodes of localized inversion
5.2. Early Miocene: Syn-extension occurring at different times during the Miocene. A first inversion in­
terval took place during the deposition of Middle to Late Miocene se­
The Early Miocene sequences (units U1 and U2) represent a syn- quences (U3 and U4) and we correlate this episode with the uplift of the
extension phase (Fig. 14). We interpret the area covered by the Dorado High. Evidence of this inversion is observed in the northern
seismic unit U1 to be the proximal domain (Fig. 12), where sediments study area near the Cuiza Fault. We interpreted this event based on the
filled moderate accommodation generated during this extension phase inversion definition in Cooper and Warren (2020) and due to: i) the
and probably during the Oligocene transtension (Gorney et al., 2007; absence of the unit U3 over the Dorado High (Figs. 5 and 9); and ii) the
Escalona and Mann, 2011; Vence and Mann, 2020). Further westward observation that units U3 and U4 are thicker over the limbs and thinner
deposition of sediments at the earliest Early Miocene did not occur, in the core of the Dorado High (Fig. 9). This differential thickness can be
probably because no accommodation was generated or due to low interpreted as indicating deposition during compression (Martínez and
sediment supply. Unit U1 was deposited in a shallow platform that Fuentes, 2022). The second episode of inversion occurred in the distal
favored the formation of carbonate banks (Escalante and thesis, 2005). offshore area (Tayrona Sub-basin) after the accumulation of Late
The top of unit U1 marks a change in the tectonic regime of the basin Miocene unit U4, probably during the deposition of the Late
supported by the downlapping termination of the overlying reflections. Miocene-Early Pliocene unit U5. Evidence for this episode includes: i)
We propose that the extension continued during the latest Early the presence of an inversion anticline affecting unit U2 and the forma­
Miocene during at which unit U2 was deposited (Fig. 14). The extension tion of footwall shortcut faults (Cooper and Warren, 2020) on the
had reached deeper margins (westward) resulting in the formation of the eastern flank of the Tayrona Sub-basin (Fig. 5); ii) units U3 and U4 are
Tayrona Depocenter (Londoño et al., 2015) where unit U2 reaches its folded over the anticline; however, here no thickness variation is
maximum thickness (Figs. 5, 7 and 10). This extension event can be observed; and iii) folding of the seismic units U2 to U4 (Fig. 10) on the
associated with transtension caused by the reactivation of Oca and Cuiza western side of the Tayrona Sub-basin. This last observation supports the
strike-slip faults (Kellogg, 1984; Gómez and thesis, 2001; Pindell and interpretation of the uplift of the South Caribbean Deformed Belt during
Kennan, 2009; Piraquive et al., 2017). Vence and Mann (2020) linked the Middle to Late Miocene that created the northwestern barrier of the
transtensional conditions during the Early Miocene to the oblique Tayrona Sub-basin (Vence and Mann, 2020).
collision of the Caribbean Plate against South America followed by the We interpret inversion deformation as linked to the reactivation of
activity of northwest-striking normal faults that allowed higher thick­ large basement-rooted faults (Fault Group 2). The reactivation of large,
ness of clastic sediment to fill the Tayrona basin. The growth of the formerly extensional faults is likely to occur since they are more sus­
thickness towards the footwall block, observed in the seismic unit U2 (i. ceptible to inversion than smaller ones (Kelly et al., 1999; Reilly et al.,
e. within the eastern flank of Tayrona Sub-basin - Fig. 10, or the 2017). Fault inversion is also controlled by the relative orientation of

19
L. Castro-Vera et al.
20

Journal of South American Earth Sciences 121 (2023) 104109


Fig. 14. Neogene and Quaternary evolution of the La Baja Guajira Basin. The figure presents and correlates the results obtained from the seismic interpretation and the basin modeling. The burial history is divided into
north and south due to the location of the modeled wells. Two global sea level scales (Hardenbol et al., 1988; Miller et al., 2020) and calculated linear sedimentation rates (LSR) are also shown.
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

compressional stress to the pre-existing faults (Panien et al., 2005; 1. Pre-extension: the acoustic basement is interpreted as a pre-
Withjack et al., 2010). Local inversion can be the consequence of extension sequence being affected by several clusters of normal
shallow subduction of the Caribbean crust beneath the South American faults (Fault Group 1). The configuration of the basement is an
Plate that affected the study area and surrounding areas (Kellogg, 1984; important controlling factor of the sedimentation in the study area.
van der Hilst and Mann, 1994; Taboada et al., 2000; Mann et al., 2006; The extension event that affected the basement is correlated with an
Gorney et al., 2007; Vence and Mann, 2020). Another regional event Oligocene transtension phase caused by right-lateral strike-slip
that might have contributed to deformation is the collision of the Pan­ faulting interpreted in previous studies. This transtension resulted in
ama Arc with western South America (Bernal-Olaya et al., 2015; León the formation of basins and horst and graben structures (Dorado
et al., 2017; Montes et al., 2019). High and Chuchupa Platform).
The end of the Middle Miocene sequence in this study represents an 2. Early Miocene syn-extension: moderate accommodation space was
unconformity that is more evident along today’s offshore La Baja Guajira generated during the earliest Early Miocene, which was filled by
Basin, and that has been previously interpreted (i.e., Audemard, 2001; sediments deposited on a shallow platform where carbonate banks
Maceralli, 1995; Gorney et al., 2007; Vence and Mann, 2020). This formed. The continuation of the extension phase in the latest Early
stratigraphic gap might indicate a change in tectonics (Audemard, 2001) Miocene reached deeper marginal areas and formed the Tayrona
related to inversion (Gorney et al., 2007) contemporaneous to and Depocenter. We assume the extension to be associated with trans­
probably triggering a forced regression (Fig. 14). tension due to the reactivation of Oca and Cuiza strike-slip faults, as
the result of the oblique collision of the Caribbean Plate against
5.4. Early Pliocene-Pleistocene: post-extension South America.
3. Middle to Late Miocene post-extension and localized inversion:
Units U5 to U7 were deposited during the Early Pliocene to Pleisto­ Sediments of units U3 and U4 filled significant residual accommo­
cene. The clastic clinoforms continuously advanced westward and were dation space created during the Early Miocene, and clinoforms
already evident in the near offshore area. We recognize submarine started prograding from east to west. Additional sediment input from
canyons/channels incising the Early Pliocene to Pleistocene sequences the south, sourced by the exhumation of adjacent mountain belts,
along the deep-water setting, which can be associated with deltaic sys­ resulted in higher sedimentation rates, the first peak of rapid tectonic
tems prograding to the north (i.e., proto-Magdalena and proto-Ran­ subsidence, and led strata to be subjected to maximum temperatures
chería rivers) that developed since the Middle Miocene and were (Well-1, Well-3, and Well-4). A forced regression event occurred at
transporting sediments from the surrounding mountain belts (Escalona the end of the Middle Miocene. Inverted reactivation of larger
and Mann, 2004; Guzman and Fisher, 2006; Escalona and Mann, 2011; basement-rooted faults (Fault Group 2) locally affected today’s
Romero et al., 2019). Vence and Mann (2020) attributed the formation offshore areas until the Late Miocene. This is attributed to the
of submarine canyons in the same area, during Early Pliocene times, to regional compressive regime as a consequence of the subduction of
an increase in the sediment influx and eustasy. These previous in­ the Caribbean Plate underneath the South American Plate and/or the
terpretations can also be related to the unconformity that characterizes collision of the Panama arc against western South America.
the end of the Pliocene sequence in this study, under which older 4. Early Pliocene to Pleistocene post-extension: a second exhumation
Miocene events are truncated (Fig. 5). event of the Andes affected the study area during the Pliocene. From
The Pliocene marks a period of significant increment of the sedi­ the Early Pliocene, localized inversion had stopped and erosive fea­
mentary input and a second peak of rapid tectonic subsidence (Fig. 14; tures (e.g. submarine canyons) formed in the deep-water of the La
Vence and Mann, 2020). These observations agree with the uplift event Baja Guajira Basin.
of the Sierra de Perijá (western border of the Maracaibo Basin) and the
final exhumation of the Andes during the Late Miocene–Pliocene (Kel­ With respect to the hydrocarbon potential, the high overall sedi­
logg, 1984; Mann et al., 2006; Parra et al., 2009; Escalona and Mann, mentation rates (up to ca. 900 m/my) calculated for the La Baja Guajira
2011; Mora et al., 2013). Basin can be interpreted as having considerably contributed to the
We assume the Pleistocene as a rather tectonically quiet period generation of microbial gas. Significant Neogene to recent tectonic ac­
supported by most of the flat burial curves of the modeled wells (Figs. 13 tivity has furthermore supported the formation of structural traps.
and 14) and no evidence of fault activity during the deposition of units
U6 and U7. The burial history at Well-6 displays continuous subsidence Funding
from the Pleistocene to recent, and the thermal history at Well-1 in­
dicates the peak of maximum temperature reached in recent times This work was supported by the following Colombian institutions:
(Figs. 13 and 14). We suggest that ongoing hydrocarbon generation can Ministerio de Educación Nacional (MEN); Ministerio de Ciencia, Tec­
be expected in today’s offshore south La Baja Guajira Basin. However, nología e Innovación (COLCIENCIAS); and Instituto Colombiano de
there is a high risk of charging traps located in onshore areas with Estudios Técnicos en el Exterior (ICETEX) (grant within the framework
thermogenic hydrocarbons generated in offshore depocenters in the of the program Colombia Científica - Pasaporte a la Ciencia, 2018).
southern La Baja Guajira Basin, since a long-distance migration would
be required. CRediT authorship contribution statement
The origin of hydrocarbons in the La Baja Guajira Basin has been
interpreted as involving microbial origin (Rice and Claypool, 1981; Katz Leidy Castro-Vera: Writing – original draft, Visualization, Software,
and Williams, 2003; Escalante, 2006). High sedimentation rates can Methodology, Investigation, Funding acquisition, Conceptualization.
trigger the generation of microbial gas (Katz, 2011), which can also be Ralf Littke: Writing – review & editing, Validation, Supervision, Soft­
assumed for the La Baja Guajira Basin (Table 4). Such conditions also ware. Stefan Back: Writing – review & editing, Validation, Supervision,
lead to high chances of preservation of possible accumulations (Qiao Software. Rocío Bernal-Olaya: Writing – review & editing, Supervision,
et al., 2021). Software, Methodology, Conceptualization. Martin Reyes: Writing –
review & editing, Software. J. Alejandro Mora-Bohórquez: Data
6. Conclusions curation, Writing - review & editing. A. Johamna Romero: Data
curation, Writing - review & editing.
The following phases can be differentiated in the La Baja Guajira
Basin on the basis of our interpretation of the seismic reflection profiles
integrated with borehole data, and constrained with 1D basin modeling:

21
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Declaration of competing interest Cardona, A., Weber, M., Wilson, R., Cordani, U., Muñoz, C.M., Paniagua, F., 2007.
Evolución tectono-magamtica de las rocas máficas-ultramáficas del Cabo de la Vela y
el Stock de Parashi, Península de la Guajira: Registro de la evolución orogénica
The authors declare that they have no known competing financial Ctretácica-Eocena del norte de Suramérica y el Caribe: XI Congreso Colombiano de
interests or personal relationships that could have appeared to influence Geología, vol. 7.
the work reported in this paper. Case, J.E., Holcombe, T.L., Martin, R.G., 1984. Map of geologic provinces in the
Caribbean region. In: Bonini, W., Hargraves, R., Shagan, R. (Eds.), The Caribbean-
South American plate boundary and regional tectonics, 162, pp. 1–30. Geological
Data availability Society of America Special Memoir.
Castillo, M.V., Mann, P., 2006. Cretaceous to Holocene structural and stratigraphic
development in south Lake Maracaibo, Venezuela, inferred from well and three-
The data that has been used is confidential. dimensional seismic data. AAPG (Am. Assoc. Pet. Geol.) Bull. 90 (4), 529–565.
Catuneanu, O., Abreu, V., Bhattacharya, J.P., Blum, M.D., Dalrymple, R.W., Eriksson, P.
Acknowledgments G., Fielding, C.R., Fisher, W.L., Galloway, W.E., Gibling, M.R., Giles, K.A.,
Holbrook, J.M., Jordan, R., Kendall, CG. St C., Macurda, B., Martinsen, O.J., Miall, A.
D., Neal, J.E., Nummedal, D., Pomar, L., Posamentier, H.W., Pratt, B.R., Sarg, J.F.,
Special thanks to Hocol S.A for providing the data used in this study. Shanley, K.W., Steel, R.J., Strasser, A., Tucker, M.E., Winker, C., 2009. Towards the
We thank Schlumberger for providing Petrel licenses for seismic and standardization of sequence stratigraphy. Earth Sci. Rev. 92, 1–33.
Cediel, F., Aguilera, R., 2011. Petroleum Geology of Colombia, Guajira and Cayos Basins,
well interpretation, and PetroMod licenses for basin modeling. vol. 8. Fondo Editorial Universidad Eafit, Medellín.
Furthermore, we thank reviewers Lucía Torrado, Fernando Martínez, Cerón-Abril, J., 2008. Crustal Structure of the Colombian Caribbean Basin and Margins.
and Luis Carlos Carvajal-Arenas for their constructive and detailed Ph.D. Dissertation, University of South Carolina, Columbia, p. 165.
Cooper, M., Warren, M.J., 2020. Inverted fault systems and inversion tectonic settings.
comments, which helps us to improve our manuscript. In: Scarselli, N., Adam, J., Chiarella, D., Roberts, D.G., Bally, A.W. (Eds.), Regional
Geology and Tectonics: Principles of Geologic Analysis. Elsevier, Amsterdam, the
References Netherlands, pp. 169–204. https://doi.org/10.1016/B978-0-444-64134-2.00009-2.
Duerto, L., Escalona, A., Mann, P., 2006. Deep structure of the Merida andes and Sierra
de Perija mountain fronts, Maracaibo Basin, Venezuela. AAPG (Am. Assoc. Pet.
Alberding, H., 1957. Application of Principles of Wrench-Fault Tectonics of Moody and
Geol.) Bull. 90, 505–528.
Hill to Northern South America, vol. 68. The Geological Society of America,
Duque-Caro, H., Reyes, R., 1999a. Seismic Biostratigraphy Study of the Guajira Region
pp. 786–790.
(Onshore and Offshore) (Dique Caro & Cia Ltda. Internal report Hocol S.A).
Allen, P.A., Allen, J.R., 2013. Basin Analysis: Principles and Application to Petroleum
Duque-Caro, H., Reyes, R., 1999b. Appendix 1. Enclosure 9 - Biostratigraphic Summary
Play Assessment, third ed. Wiley-Blackwell, Berlin, ISBN 978-0-470-67377-5.
Chuchupa-1. Internal report Hocol S.A.
Alvarez, W., 1967. Geology of the Simarua and Carpintero Areas. University of
Duque-Caro, H., Reyes, R., 2000. Appendix 1. Enclosure 3 - Biostratigraphy Summary
Princeton, Guajira Peninsula, Colombia, p. 164.
Aruchara-1. Duque Caro & Cia Ltda. Internal report Hocol S.A.
Alvarez, W., 1971. Fragmented Andean Belt of Northern Colombia. The Geological
Ecopetrol, 1988. Geological Report Tinka-1. Internal Report.
Society of America, pp. 77–96. Memoir 130.
Escalante, C.E., 2006. Seismic Stratigraphy of the Tayrona Depression-Offshore Baja
Atlas International, Western, 1990. Descripción Litológica Detalladas Y Análisis
Guajira Basin, Colombia. Asociación Colombiana de Geólogos y Geofísicos del
Petrográfico Detallado, Reporte Final. Internal Report Ecopetrol.
Petróleo (ACGGP).
Audemard, F.A., 1993. Neótectonique, Sismotectonique et Alea Sismique du Nord-ouest
Escalante, C.E., 2005. Integrated Seismic Stratigraphic and 1-D Basin Analysis of the
du Vénézuéla (Systéme de failles de Oca-Ancón). PhD thesis. Universite Montpellier
Tayrona Depression: Offshore Baja Guajira Basin. University of Oklahoma,
II, p. 369.
Colombia, p. 134. M.S. thesis.
Audemard, F.A., 1995. La cuenca Terciaria de Falcón, Venezuela noroccidental: Síntesis
Escalona, A., Mann, P., Bolivar Group, 2004. Tectonic Reconstructions of Sedimentary
estratigráfica, génesis e inversión tectónica: IX Congreso Latinoamericano Geológico.
Basins Associated with the Proto-Maracaibo and Proto-Orinoco Rivers: New
Audemard, F.A., 1996. Paleoseismicity studies on the Oca-Ancon fault system,
Constraints from Bolivar and Gulfrex Seismic Data (Abs. American Geophysical
northwestern Venezuela. Tecotonophysics 259, 67–80.
Union Fall Meeting, San Francisco, CD-ROM.
Audemard, F.A., 2001. Quaternary tectonics and present stress tensor of the inverted
Escalona, A., Mann, P., 2011. Tectonics, basin subsidence mechanisms, and
northern Falcón Basin, northwestern Venezuela. J. Struct. Geol. 23, 431–453.
paleogeography of the Caribbean-South American plate boundary zone. Mar. Petrol.
Audemard, F.A., 2009. Key issues on the post-mesozoic southern Caribbean Plate
Geol. 28, 8–9.
boundary. Geol. Soc., London, Spec. Publ. 328 (1), 569–586.
Feo-Codecido, G., 1972. Breves Ideas Sobre la Estructura de la Falla de Oca. VI
Audemard, F.A., Singer, A., Rodríguez, J.A., Beltrán, C., 1994. Definición de la traza
Conferencia Geológica del Caribe, Venezuela, 184–190.
activa del sistema de fallas de Oca–Ancón, Noroccidente de Venezuela. VII Congr.
García, M., Cruz, L.E., Mier, R., Vásquez, M., Jiménez, M., Moreno, M., 2008. Evolución
Venezolano Geofısica, Caracas, Venezuela, pp. 43–51.
térmica de la subcuenca de la Baja Guajira. Informe final UIS-ANH.
Avé Lallemant, H., 1997. Transpression, displacement partitioning, and exhumation in
General Bathymetric Chart of Oceans. In: 2020GEBCO gridded bathymetry data
the eastern Caribbean/South American plate boundary zone. Tectonics 16, 272–289.
download, Retrieved November, 2020 from. https://download.gebco.net/.
Barrero, D., Pardo, A., Vargas, C., Martinez, J., 2007. Colombian Sedimentary Basins:
Global Mapper Online Data Source, 2020. ASTER GDEM v3 Worldwide Elevation Data.
Nomenclature, Boundaries and Petroleum Geology, a New Proposal. ANH, Bogotá.
Gomez, I., 2001. Structural style and evolution of the Cuisa fault system, Guajira,
Bayona, G., Ochoa, F.L., Cardona, A., Jaramillo, C., Montes, C., Tchegliakova, N., 2007.
Colombia. M.Sc. Thesis. University of Houston 147 p.
Procesos orogénicos del Paleoceno para la cuenca de Ranchería (Guajira, Colombia)
Gómez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., Kelley, S., 2005. Syntectonic
y áreas adyacentes definidos por análisis de procedencia. Geol. Colomb. 32, 21–46.
cenozoic sedimentation in the northern middle magdalena valley basin of Colombia
Bermúdez, M.A., Bernet, M., Kohn, B.P., Brichau, S., 2019. In: Cediel, F., Shaw, R.P.
and implications for exhumation of the northern andes. Geol. Soc. Am. Bull. 117,
(Eds.), Exhumation-Denudation History of the Maracaibo Block, Northwestern South
547–569. https://doi.org/10.1130/B25454.1.
America: Insights from Thermochronology. https://doi.org/10.1007/978-3-319-
Gómez, J., Montes, N.E., Nivia, A., Diederix, H., compiladores, 2015. Mapa Geológico de
76132-9_13. Geology and Tectonics of Northwestern South America, Frontiers in
Colombia 2015. Servicio Geológico Colombiano, Bogotá. https://doi.org/10.32685/
Earth Sciences.
10.143.2015.935. Escala 1:1 000 000.
Bernal-Olaya, R., Mann, P., Escalona, A., 2015. Cenozoic tectonostratigraphic evolution
Gorney, D., Escalona, A., Mann, P., Magnani, M.B., BOLIVAR Study Group, 2007.
of the Lower Magdalena Basin, Colombia: an example of an under- to over filled
Chronology of Cenozoic tectonic events in western Venezuela and the Leeward
forearc basin. In: Bartolini, C., Mann, P. (Eds.), Petroleum geology and potential of
Antilles based on integration of offshore seismic reflection data and on-land geology.
the Colombian Caribbean margin, 108, pp. 345–398. AAPG Memoir.
AAPG (Am. Assoc. Pet. Geol.) Bull. 91 (5), 653–684.
Blanco, J.M., Mann, P., Nguyen, L., 2015. Location of the Suture Zone separating the
Guzman, J., Fisher, W., 2006. Early and middle Miocene depositional history of the
Great Arc of the Caribbean from continental crust of northwestern South America
Maracaibo Basin, western Venezuela. AAPG (Am. Assoc. Pet. Geol.) Bull. 90,
inferred from regional gravity and magnetic data. In: Bartolini, C., Mann, P. (Eds.),
625–656.
Petroleum Geology and Potential of the Colombian Caribbean Margin: AAPG
Hantschel, T., Kauerauf, A.I., 2009. Fundamentals of Basin and Petroleum Systems
Memoir 108, pp. 161–178.
Modeling. Springer, Dordrecht.
Bruns, B., Di Primo, R., Berner, U., Littke, R., 2013. Petroleum system evolution in the
Hardenbol, J., Haq, B.U., Vail, P.R., 1988. Mesozoic and Cenozoic Chronostratigraphy
inverted Lower Saxony Basin, northwest Germany: a 3D basin modeling study.
and Cycles of Sea-Level Change (Exxon Production Research Company).
Geofluids 13, 246–271.
Hosie, A., 1994. Seismic Stratigraphy of the Southern Offshore Guajira Basin. University
Burke, K., 1988. Tectonic evolution of the caribbean. Annu. Rev. Earth Planet Sci. 16,
of Oklahoma, Colombia, p. 141. Norman.
201–230. https://doi.org/10.1146/annurev.ea.16.050188.001221.
Instituto Colombiano del Petróleo, División de exploración y Producción Área
Burke, K., Cooper, C., Dewey, J., Mann, P., Pindell, J., 1984. Caribbean tectonics and
Geoquímica, 2000. Evaluación geoquímica de pozos Mero-1 y San José-1. Internal
relative plate motions. In: Bonini, W., Hargraves, R., Shagam, R. (Eds.), The
report Hocol. S.A.
Caribbean-South American Plate Boundary and Regional Tectonics. Geological
Katz, B.J., 2001. Lacustrine basin hydrocarbon exploration-current thoughts.
Society of America, pp. 31–57.
J. Paleolimnol. 26, 161–179.
Burnham, A.K., 2016. Evaluation of alternatives to Easy%Ro for calibration of basin and
Katz, B., 2011. Microbial processes and natural gas accumulations. Open Geol. J. 5,
petroleum system models: Extended abstract. EAGE, Paris, France.
75–83.
Burnham, A.K., 2019. Kinetic models of vitrinite, kerogen, and bitumen reflectance.
Organic Geochemistry, 131. p. 50–59.

22
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Katz, B., Williams, K., 2003. Biogenic gas potential offshore Guajíra peninsula, Colombia. Lorente, M.A., Pindell, J.L. (Eds.), The origin and evolution of the Caribbean plate,
In: Bartolini, C., Buffler, R.T., Blickwede, J. (Eds.), The Circum-Gulf of Mexico and 328, pp. 1–55. GSL Special Publications.
the Caribbean: Hydrocarbon Hábitats, Basin Formation, and Plate Tectonics: AAPG Pindell, J.L., Cande, S.C., Pitman, W.C., Rowley, D.B., Dewey, J.F., Labrecque, J.,
Memoir 79, pp. 173–175. Haxby, W., 1988. A plate-kinematics framework for models of Caribbean evolution.
Kellogg, J.N., 1984. Cenozoic tectonic history of the Sierra de Perijá, Venezuela- Tectonophysics 155, 121–138.
Colombia, and adjacent basins. In: Bonini, W.E., Hargraves, R.B., Shagam, R. (Eds.), Pindell, J.L., Higgs, R., Dewey, J.F., 1998. Cenozoic palinspastic reconstruction,
The Caribbean-South American plate boundary and regional tectonics, 162, paleogeographic evolution and hydrocarbon setting of the northern margin of South
pp. 239–261. GSA Memoir. America. In: Pindell, J.L., Drake, C. (Eds.), Paleogeographic Evolution and Non-
Kelly, P.G., Peacock, D.C.P., Sanderson, D.J., McGuirk, A.C., 1999. Selective reverse glacial Eustacy, pp. 45–85. North America: Society for Sedimentary Geology SEPM
reactivation of normal faults, and deformation around reverse reactivated faults in Special Publication.
the Mesozoic of the Somerset coast. J. Struct. Geol. 21, 493509. Pindell, J., Kennan, L., Stanek, K.P., Maresch, W.V., Draper, G., 2006. Foundations of
Lawrence, L.J., 1958. Geological Report Saure-1. Internal Report Hocol S.A. Gulf of Mexico and caribbean evolution: eight controversies resolved. Geol. Acta 4
León, S., Cardona, A., Parra, M., Sobel, E.R., Jaramillo, J.S., Glodny, J., Valencia, V.A., (1–2), 303–341.
Chew, D., Montes, C., Posada, G., Monsalve, G., Pardo-Trujillo, A., 2017. Transition Piraquive, A., Pinzón, E., Kammer, A., Bernet, M., von Quadt, A., 2017. Early Neogene
from collisional to subduction-related regimes: an example from Neogene Panama- unroofing of the Sierra Nevada de Santa Marta, as determined from detrital
Nazca-South America interactions. Tectonics 37, 119–139. https://doi.org/10.1002/ geothermochronology and the petrology of clastic basin sediments. Geological
2017TC004785. Society of America Bulletin.
Londoño, J., Schiek, C., Biegert, E., 2015. Basement architecture of the southern Qiao, J., Grohmann, S., Baniasad, A., Zhang, C., Jiang, Z., Littke, R., 2021. High
caribbean basin, Guajira offshore, Colombia. In: Bartolini, C., Mann, P. (Eds.), microbial gas potential of Pleistocene lacustrine deposits in the central Qaidam
Petroleum Geology and Potential of the Colombian Caribbean Margin: AAPG Basin, China: an organic geochemical and petrographic assessment. Int. J. Coal Geol.
Memoir 108, pp. 85–102. 245 (1 September 2021), 103818.
Lugo, J., Mann, P., 1995. Jurassic-eocene tectonic evolution Maracaibo Basin, Venezuela. Raasveldt, H.C., 1956. Fallas de rumbo en el nordeste de Colombia. Revista de Petróleo
In: Tankard, A., Suarez, S., Welsink, H. (Eds.), Petroleum Basins of South America: 7, 19–26.
AAPG Memoir 62, pp. 699–725. Ramirez, V., 2007. Stratigraphic Framework and Petroleum Systems Modeling, Guajira
MacDonald, W., 1964. Geology of the Serrania de Macuira area Guajira Peninsula. Basin Northern Colombia. M.S. thesis. University of Alabama.
Unpublished Ph.D. dissertation, Princeton University, Colombia, p. 167. Ramirez, V., Vargas, L.S., Rubio, C., Nino, H., Mantilla, O., 2015. Petroleum systems of
MacDonald, W., Opdyke, N., 1972. Tectonic rotations suggested by paleomagnetic results the Guajira Basin, northern Colombia. In: Bartolini, C., Mann, P. (Eds.), Petroleum
from northern Colombia, south America. J. Geophys. Res. 77, 5720–5730. Geology and Potential of the Colombian Caribbean Margin: AAPG Memoir 108,
Maceralli, C.E., 1995. Cenozoic sedimentation and tectonics of the southwestern pp. 399–430.
caribbean pull-apart basin, Venezuela and Colombia. In: Tankard, A.J., Suárez, R.S., Rangel, A., Katz, B., Ramirez, V., Vaz dos Santos Neto, E., 2003. Alternative
Welsink, H.J. (Eds.), Petroleum Basins of South America: AAPG Memoir 62, interpretations as to the origin of the hydrocarbons of the Guajira Basin, Colombia.
pp. 757–780. Mar. Petrol. Geol. 20, 129–139.
Mann, P., 1999. Caribbean sedimentary basins: classification and tectonic setting from Reilly, F.A., 1958. Final Geological Report Guaitapa-1. Internal Report Hocol S.A.
Jurassic to present. In: Mann, P. (Ed.), Caribbean basins, sedimentary basins of the Reilly, C., Nicol, A., Walsh, J., 2017. Importance of pre-existing fault size for the
world, 4, pp. 3–31. Elsevier Science. evolution of an inverted fault system. In: Childs, C., Holdsworth, R.E., Jackson, C.A.-
Mann, P., Schubert, C., Burke, K., 1990. Review of caribbean neotectonics. In: Dengo, G., L., Manzocchi, T., Walsh, J.J., Yielding, G. (Eds.), The Geometry and Growth of
Case, J. (Eds.), The Caribbean Region, pp. 307–338. Geological Society of America, Normal Faults, 439, pp. 447–463. https://doi.org/10.1144/SP439.2. Geological
The Geology of North America, vol. H. Society, London, Special Publications, London.
Mann, P., Escalona, A., Castillo, M.V., 2006. Regional geologic and tectonic setting of the Renz, O., 1956. Cretaceous in Western Venezuela and the Guajira (Colombia). 20 Congr.
Maracaibo supergiant basin, western Venezuela. AAPG (Am. Assoc. Pet. Geol.) Bull. Geol. Internat., México.
90 (4), 445–477. Renz, O., 1960. Geología de la parte Sureste de la Península de la Guajira (República de
Martínez, F., Fuentes, G., 2022. Inverted Structures in the Western Central Andes Thrust Colombia). III Congreso Geológico Venezolano, pp. 317–347.
Belt Front. Andean Structural Styles. https://doi.org/10.1016/B978-0-323-85175- Rice, D., Claypool, G., 1981. Generation, accumulation, and resource potential of
6.00037-7 (Chapter 37). biogenic gas. AAPG (Am. Assoc. Pet. Geol.) Bull. 65, 5–25.
Miller, K.G., Browning, J.V., Schmelz, W.J., Kopp, R.E., Mountain, G.S., Wright, J.D., Rod, E., 1956. Strike-slip faults of northern Venezuela. AAPG (Am. Assoc. Pet. Geol.)
2020. Smoothed Cenozoic sea-level relative to modern from deep-sea geochemical Bull. 40, 457–476. https://doi.org/10.1306/5CEAE3E5-16BB-11D7-
and continental margin records. PANGAEA. https://doi.org/10.1594/ 8645000102C1865D.
PANGAEA.923139. Rodon, S., Littke, R., 2005. Thermal maturity in the Central European Basin system
Montes, C., Hatcher, R.D., Restrepo-Pace, P.A., 2005. Tectonic reconstruction of the (Schleswig-Holstein area): results of 1D basin modelling and new maturity maps. Int.
northern Andean blocks: oblique convergence and rotations derived from the J. Earth Sci. 94, 815–833. https://doi.org/10.1007/s00531-005-0006-1.
kinematics of the Piedras-Girardot area, Colombia. Tectonophysics 399, 221–250. Rollins, J.F., 1965. Stratigraphy and Structure of the Guajira Peninsula, Northwestern
Montes, C., Rodriguez-Corcho, A.F., Bayona, G., Hoyos, N., Zapata, S., Cardona, A., 2019. Venezuela and Northeastern Colombia. University of Nebraska, Lincoln (Nebraska).
Continental margin response to multiple arc-continent collisions: the northern Romero, A.J., Mora, J.A., Vélez, V., 2019. Controls on the Miocene to recent infill of the
Andes-Caribbean margin. Earth Sci. Rev. 198, 102903. lower Guajira Basin of northern Colombia. In: Poster Presented at the AAPG
Mora, A., Reyes-Harker, A., Rodríguez, G., Tesón, E., Ramírez-Arias, J.C., Parra, M., International Conference and Exhibition, Buenos Aires, Argentina, vol. 27, p. 30.
Caballero, V., Mora, J.P., Quintero, I., Valencia, V., Ibáñez, M., Horton, B.K., August,.
Stockli, D.F., 2013. In: Nemcok, M., Mora, A., Cosgrove, J.W. (Eds.), Inversion Sanchez, J., Mann, P., Carvajal-Arenas, L.C., Bernal-Olaya, R., 2019. Regional transect
Tectonics under Increasing Rates of Shortening and Sedimentation: Cenozoic across the western Caribbean Sea based on integration of geologic, seismic
Example from the Eastern Cordillera of Colombia, Thick-Skin-Dominated Orogens: reflection, gravity, and magnetic data. AAPG (Am. Assoc. Pet. Geol.) Bull. 103 (2),
from Initial Inversion to Full Accretion, vol. 377. Geological Society, London, Special 303–343.
Publications, pp. 411–442. Senglaub, Y., Littke, R., Brix, M.R., 2005. Numerical modelling of burial and temperature
Mora-Bohórquez, J.A., Ibánez-Mejia, M., Oncken, O., de Freitas, M., Vélez, V., Mesa, A., history as an approach for an alternative interpretation of the Bramsche anomaly,
Serna, L., 2017. Structure and age of the Lower Magdalena Valley basin basement, Lower Saxony Basin. Int. J. Earth Sci. https://doi.org/10.1007/s00531-005-0033-y.
northern Colombia: new reflection-seismic and U-Pb-Hf insights into the termination Soulas, J.P., Giraldo, C., Bonnot, D., Lugo, M., 1987. Actividad cuatemaria y
of the central andes against the Caribbean basin. J. S. Am. Earth Sci. 74, 1–26. caracterlsticas sismogénicas del sistema de fallas de Oca-Ancón y de las fallas de
https://doi.org/10.1016/j.jsames.2017.01.001. Lagarto, Urumaco, Rio Seco y Pedregal. Afinamiento de las características
Panien, M., Schreurs, G., Pfiffner, A., 2005. Sandbox experiments on basin inversion: sismogénicas de las fallas de Mene Grande y Valera. Co. Rep., MARAVEN SA,
testing the influence of basin orientation and basin fill. J. Struct. Geol. 27, 433–445. Caracas.
Parra, M., Mora, A., Sobel, E.R., Strecker, M.R., González, R., 2009. Episodic orogenic Sweeney, J.J., Burnham, A.K., 1990. Evaluation of a simple model of vitrinite reflectance
front migration in the northern Andes: constraints from low-temperature based on chemical kinetics. AAPG Bull. 74, 1559–1570.
thermochronology in the Eastern Cordillera, Colombia. Tectonics 8, TC4004. Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J.,
https://doi.org/10.1029/2008TC002423. Rivera, C., 2000. Geodynamics of the Northern Andes: subductions and
Patiño, A.M., Parra, M., Ramírez, J.C., Sober, E.R., Glodny, J., Almendral, A., intracontinental deformation (Colombia). Tectonics 19, 787.
Echeverri, S., 2019. Thermochronological constraints on Cenozoic exhumation along Texas Petroleum Company, 1989. Weekely Geological Report Merluza-1. Internal report
the southern Caribbean: the Santa Marta range, northern Colombia. In: Andean Hocol S.A.
Tectonics. https://doi.org/10.1016/B978-0-12-816009-1.00007-1. Tschanz, C.M., Marvin, R.F., Cruz-B, J., Menhert, H., Cebula, G.T., 1974. Geologic
Perez, V., Sutherland, J., Mora, C., 1988. Análisis de sections degadas Tinka-1. Internal evolution of the Sierra Nevada de Santa Marta, northeastern Colombia. Geol. Soc.
report Hocol S.A. Am. Bull. 85, 273–284.
Pindell, J., 1993. Mesozoic-Cenozoic paleogeographic evolution of northern South van der Hilst, R., Mann, P., 1994. Tectonic implications of tomographic images of
America. AAPG Bulleting (Am. Assoc. Pet. Geol.) vol. 77 no CONF-930306-. subducted lithosphere beneath northwestern South America. Geology 22, 451–454.
Pindell, J., Barrett, S., 1990. Geological evolution of the Caribbean region: a plate Vargas, C.A., Montes, L.A., Ortega, C., 2012. Geología estructural y estratigrafía del área
tectonic perspective. In: Dengo, G., Case, J. (Eds.), The Caribbean Region. Geological Majayura (Guajira). Revista de la Academia Colombiana de Ciencias Exactas, Físicas
Society of America, The Geology of North America, v. H, pp. 405–432. y Naturales, 36(140), p. 385–398.
Pindell, L.J., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and Vence, E., 2008. Subsurface Structure, Stratigraphy, and Regional Tectonic Controls of
northern South America in the mantle reference frame: an update. In: James, K.H., the Guajira Margin of Northern Colombia. M.Sc. thesis, Austin, Texas, p. 128.

23
L. Castro-Vera et al. Journal of South American Earth Sciences 121 (2023) 104109

Vence, E., Mann, P., 2020. Subsurface basement, structure, stratigraphy, and timing of Withjack, M., Baum, M.S., Schlische, R.W., 2010. Influence of pre-existing fault fabric on
regional tectonic events affecting the Guajira margin of northern Colombia. In: inversion- related deformation: a case study of the inverted Fundy rift basin,
Interpretation, vol. 8, p. 4. November 2020); p. ST69–ST105. southeastern Canada. Tectonics 29, 1–22.
Villagómez, D., Spikings, R., Mora, A., Guzmán, G., Ojeda, G., Cortés, E., van der Lelij, R., Wygrala, B.P., 1989. Integrated study of an oil field in the southern Po basin, northern
2011. Vertical tectonics at a continental crust-oceanic plateau plate boundary zone: Italy. In: Jiil-Rep 2313, Research Centre Jiilich. Dissertation, Koln University, Jiilich,
fission track thermochronology of the Sierra Nevada de Santa Marta, Colombia. 0366-0885.
Tectonics 30, TC4004. https://doi.org/10.1029/2010TC002835. Zuluaga, C.A., Pinilla, A., Mann, P., 2015. Jurassic silicic volcanism and associated
Villamil, T., 1999. Campanian-Miocene tectonostratigraphy, depocenter evolution and continental-arc basin in northwestern Colombia (southern boundary of the
basin development of Colombia and western Venezuela. Palaeogeogr. Caribbean plate). In: Memoir 108: Petroleum Geology and Potential of the
Palaeoclimatol. Palaeoecol. 153, 239–275. Colombian Caribbean Margin, pp. 137–160. https://doi.org/10.1306/
13531934M1083640.

24

You might also like