You are on page 1of 26

EAGE

Basin Research (2015) 1–26, doi: 10.1111/bre.12154

The detrital record of late-Miocene to Pliocene


surface uplift and exhumation of theVenezuelan
Andes in the Maracaibo and Barinas foreland basins
M. A. Berm udez,*,† C. Hoorn,‡ M. Bernet,† E. Carrillo,§,1 P. A. van der Beek,† J. I.
Garver,¶ J. L. Mora* and K. Mehrkian†
*Escuela de Geologıa, Minas y Geofısica, Facultad de Ingenierıa, Universidad Central de Venezuela,
Caracas, Venezuela
†Institut des Science de la Terre, Universite Joseph Fourier, Grenoble, France
‡Institute for Biodiversity and Ecosystem Dynamics (IBED), University of Amsterdam, Amsterdam,
The Netherlands
§Instituto de Ciencias de la Tierra, Facultad de Ciencias, Universidad Central de Venezuela, Caracas,
Venezuela
¶Geology Department, Union College, Schenectady, NY, USA

ABSTRACT
A multidisciplinary approach, combining sediment petrographic, palynological and thermochrono-
logical techniques, has been used to study the Miocene-Pliocene sedimentary record of the evolution
of the Venezuelan Andes. Samples from the Maracaibo (pro-wedge) and Barinas (retro-wedge) fore-
land basins, proximal to this doubly vergent mountain belt, indicate that fluvial and alluvial-fan sedi-
ments of similar composition were shed to both sides of the Venezuelan Andes. Granitic and gneissic
detritus was derived from the core of the mountain belt, whereas sedimentary cover rocks and
uplifted foreland basin sediments were recycled from its flanks. Palynological evidence from the
Maracaibo and Barinas basins constrains depositional ages of the studied sections from late Miocene
to Pliocene. The pollen assemblages from the Maracaibo Basin are indicative of mountain vegetation,
implying surface elevations of up to 3500–4000 m in the Venezuelan Andes at this time. Detrital
apatite fission-track (AFT) data were obtained from both stratigraphic sections. In samples from the
Maracaibo basin, the youngest AFT grain-age population has relatively static minimum ages of
5  2 Ma, whereas for the Barinas basin samples AFT minimum ages are 7  2 Ma. With excep-
tion of two samples collected from the Eocene Pag€ uey Formation and from the very base of the Mio-
cene Parangula Formation, no evidence for resetting and track annealing in apatite due to burial
heating in the basins was found. This is supported by rock-eval analyses on organic matter and ther-
mal modelling results. Therefore, for all other samples the detrital AFT ages reflect source area cool-
ing and impose minimum age constraints on sediment deposition. The main phase of surface uplift,
topography and relief generation, and erosional exhumation in the Venezuelan Andes occurred dur-
ing the late Miocene to Pliocene. The Neogene evolution of the Venezuelan Andes bears certain sim-
ilarities with the evolution of the Eastern Cordillera in Colombia, although they are not driven by
exactly the same underlying geodynamic processes. The progressive development of the two moun-
tain belts is seen in the context of collision of the Panama arc with northwestern South America and
the closure of the Panama seaway in Miocene times, as well as contemporaneous movement of the
Caribbean plate to the east and clock-wise rotation of the Maracaibo block.

basins and reactivated faults, driven by the interaction of


INTRODUCTION
the Nazca plate, Caribbean plate, South American plate
The Northern Andes in northwestern South America are and Panama arc (Fig. 1), and modulated by climatic con-
a complex assemblage of mountain belts, sedimentary ditions (e.g. Dengo & Covey, 1993; Colletta et al., 1997;
Mora et al., 2008; Hoorn et al., 2010; Farris et al., 2011).
Correspondence: M.A. Berm udez, Escuela de Geologıa, Minas While the evolution of the Western and Central Cordil-
y Geof́ sica, Facultad de Ingeniería, Universidad Central de leras in Colombia is related to Late Cretaceous arc-terrane
Venezuela, Caracas, Venezuela. E-mail: mauricio.bermudez@
ucv.ve or maberce@yahoo.com accretion (e.g. Case et al., 1990; Dengo & Covey, 1993;
1
Present address: Robertson CGG Tyn-y-Coed, Pentywyn Gregory-Wodzicki, 2000; Cediel et al., 2003), the driving
Road, Llandudno, LL30 1SA, UK factors of the evolution of the Eastern Cordillera in

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 1

M.A. Bermudez et al.

CARIBBEAN
15°N
CARIBBEAN PLATE

Antilles
Lesser
COCOS
–1
1–2 cm year
SOUTH AMERICA

NAZCA

VF
?

SBF
MB EB 10°
A
BF
Magdalena SM

VA
basin Venezuela
COCOS
PLATE CCC
9 cm year–1 ECC Llanos SOUTH AMERICAN PLATE
Orientales Guyana Shield
NAZCA basin 5°
PLATE
6 cm year–1
WCC
Colombia

Elevation/depth (m)
>3000
Brazil 2000
1000
0
Ecuador –1000
–2000
–3000
–4000 0°
–5000
–6000
Amazonas <–7000

basin VF Valera Fault ECC Eastern Colombian Cordillera


BF Boconó Fault zone SBF Santa Marta-Bucaramanga fault
MB Maracaibo basin VA Venezuelan Andes
EBA El Baúl Arch SM Santander Massif
WCC Western Colombian Cordillera Barinas-Apure basin km
CCC Central Colombian Cordillera 0 100200
5°S
85° 80° 75° 70° 65° 60° 55°W

Fig. 1. Major tectonic features of northern South America (after Colmenares & Zoback, 2003). Topography and bathymetry is based
on GeoMapApp developed by the IEDA database group at Columbia University’s Lamont-Doherty Earth Observatory (see http://
www.geomapapp.org) square shows location of Fig. 2.

Colombia and the Venezuelan Andes are under discus- Andes in Colombia to the NW and W and the Guyana
sion. Therefore, the uplift and exhumation histories of Shield to the SE (Fig. 1).
these two mountain belts hold important information on The pro-wedge Maracaibo foreland basin, flanking the
the tectonic evolution and possible climatic feedback in Venezuelan Andes to the north, contains Miocene-
this region (e.g. Mora et al., 2008, 2010a,b; Parra et al., Pliocene detrital deposits that are divided into the Palmar,
2009a,b, 2010; Berm udez et al., 2010, 2011, 2013). Of Isnot
u and Betijoque formations (Lorente, 1986; James,
particular interest in this respect are the timing of surface 2000). These deposits grade from shallow marine
uplift, the creation of high topography, and its relation- mudrocks to fluvial mudrocks, sandstones and conglom-
ships with the nature and timing of sedimentation and erates, derived from erosion of the center and northern
subsequent thermal history of sedimentary basins flanking flank of the uplifting mountain belt.
the Venezuelan Andes. The Barinas-Apure basin, here simply referred to as
The Venezuelan (or Merida) Andes in northwestern the Barinas basin (Figs 1 and 2), is the retro-foreland
Venezuela result from rotation and oblique convergence basin on the southern flank of the Venezuelan Andes.
between the Maracaibo continental block to the north and This basin connects with the Llanos basin in Colombia to
the South American plate to the south (Fig. 1; Case et al., the SW, and forms part of the Orinoco River drainage
1990; Colletta et al., 1997; Montes et al., 2005; Pindell & basin. The Miocene-Pliocene deposits of the Barinas
Kennan, 2009). It forms a doubly vergent mountain belt basin include the Parangula and Rıo Yuca formations.
with the pro-wedge facing northwest, related to Both formations consist of fluvial mudrocks, sandstones
underthrusting of the Maracaibo Block below South and conglomerates, derived from erosion of the center
America (Colletta et al., 1997). Convergence started in and southern flank of the Venezuelan Andes. The Mio-
the late Paleogene and resulted in major surface uplift, cene-Pliocene history of surface uplift and erosional
evolution of the drainage pattern and exhumation of the exhumation of the Venezuelan Andes is thus preserved in
Venezuelan Andes during the late Miocene (Kohn et al., these synorogenic sedimentary rocks, which makes them
1984; Hoorn et al., 1995; Berm udez et al., 2010, 2011, important for understanding the long-term evolution of
2013). The growth of the Venezuelan Andes apparently this mountain belt.
caused the separation of the Maracaibo and Barinas- The driving forces and patterns of exhumation of the
Apure basins (Guzman & Fisher, 2006), which previously Venezuelan Andes were described in a series of previous
formed a continuous basin bounded by the northern studies employing bedrock and detrital apatite fission-

© 2015 The Authors


2 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

Fig. 2. 42° clockwise rotated geologic map of the study area in the Venezuelan Andes (modified after Hackley et al., 2005), indicating
the locations of in-situ AFT data (Kohn et al., 1984; Berm udez et al., 2010, 2011), modern river AFT samples (locations from
Berm udez et al., 2013; blue stars), new synorogenic detrital AFT (red hexagons) and palynology (black circles) samples in the Mara-
caibo and Barinas basins for this study. Upper left inset shows major fault systems and delimited tectonic blocks of the Venezuelan
Andes as defined by Berm udez et al. (2010): CATB, Cerro Azul Thrust; CB, Caparo; EB, Escalante; ECB, El Carmen; SLCB, Sierra
La Culata; SNB, Sierra Nevada; and TB, Trujillo blocks. Faults abbreviations correspond to: VFS Valera fault system; BFS Bocono
fault system; CSAFS Central-Sur Andino fault system; CFS Caparo fault system; BurF Burbusay fault; NWFTB and SEFTB are
North-Western and South-Eastern Fold-Thrust Belts, respectively.

track (AFT) thermochronology (Kohn et al., 1984; data has proven difficult to obtain, mainly because these
Berm udez et al., 2010, 2011, 2013). Here, we extend these formations crop out discontinuously, are poor in fossil
studies by focusing on the Maracaibo and Barinas basin content and difficult to date with magnetostratigraphy due
sediments and incorporating complementary methodolo- to the presence of coarse-grained lithologies (Betijoque
gies, including sediment petrography, palynology and and Rıo Yuca formations), or strong secondary magnetiza-
organic maturity data. These data were collected from tion (Parangula Formation; Erikson et al., 2012). Finally,
stratigraphic sections in both the pro- and retro-wedge we discuss similarities and differences between the evolu-
foreland basins, to retrieve the exhumation record on both tion of the Venezuelan Andes and the Eastern Cordillera
sides of the Venezuelan Andes. in Colombia, in the context of the Caribbean geodynamics.
Sediment petrographic analyses on conglomerates in the
field and on sandstones in thin sections provide informa-
tion on source-rock lithologies and sediment provenance. GEOLOGY OF THE VENEZUELAN
The palynology and thermochronology data provide inde- ANDES AND ASSOCIATED FORELAND
pendent constraints on stratigraphic ages. Rock-Eval anal-
BASINS
yses of organic matter were performed to obtain
independent constraints on the thermal history of the sedi- Basin formation in northwestern South America started
mentary basins. For these, detailed chrono-stratigraphic in the late Jurassic, with development of graben systems

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 3

M.A. Bermudez et al.

during Atlantic rifting (Case et al., 1990; Pindell & deposition of the shallow marine clastic La Rosa/Palmar
Kennan, 2009). These grabens were filled with clastic sed- Formation followed by the Upper Miocene fluvial Isnot u
imentary rocks of the La Quinta Formation. Continuous Formation and the Pliocene alluvial fan deposits of the
rifting led to development of a passive margin during the Betijoque Formation. The Betijoque Formation can be
Early Cretaceous, in which the La Luna Formation, a further divided into the Vich u and Sanalejos members
calcareous shale unit that is the source rock of 98% of all (Guerrero et al., 2009). Proximal parts of the Palmar,
petroleum reserves in the Maracaibo basin (Escalona & Isnot
u and Betijoque formations have been tilted to the
Mann, 2006), was deposited. The La Luna Formation is north and exposed along the northern flank of the
overlain by the Upper Cretaceous Colon Formation and Venezuelan Andes since the late Pliocene-Pleistocene
the Paleocene Guasare Formation. The tectonic setting of (Figs 2 and 3a). In total, the pro-wedge foreland basin of
the Maracaibo basin changed from a passive margin to a the Venezuelan Andes reaches a maximum thickness of
foreland basin in the late Paleocene, due to the onset of about 6 km (De Toni & Kellogg, 1993).
oblique convergence between the Caribbean and South Miocene-Pliocene flexural subsidence was caused by
American plates (Lugo & Mann, 1995). Clastic sediments, crustal loading of the rising Venezuelan Andes, which
derived from the uplifted and eroded northern Colombian resulted from oblique collision of the Maracaibo block
Andes, were transported into the Maracaibo basin, result- with the South American plate (De Toni & Kellogg,
ing in deposition of the Eocene shallow-marine/deltaic 1993). This change in regional plate kinematics may have
Misoa Formation and migration of the basin depocenter been driven by collision of the Panama arc with the north-
to the west. The mainly quartz-arenitic Misoa Formation ern Andes in Colombia (Kohn et al., 1984; Colletta et al.,
is one of the two main reservoir rocks of the Maracaibo 1997), resulting in clockwise rotation of the Maracaibo
basin today (Escalona & Mann, 2006). The reservoir is block and the Santa Marta massif (Fig. 1; Montes et al.,
sealed by low-permeability sediments on top of a regional 2005, 2010).
Late-Eocene/Oligocene unconformity, which indicates The Barinas basin is limited to the southeast by the
uplift and partial erosion of the basin deposits during the Guyana shield, to the west by the Eastern Cordillera of
Oligocene. Subsidence recommenced in the southern the Colombian Andes, and to the northeast by the El Ba ul
Maracaibo basin in the Early-Middle Miocene, with Arch (Fig. 1). During the Late Cretaceous and Eocene,

Cerro Alto
NW Sabana de Betijoque Isnotú SE
Mendoza Betijoque Fm Isnotú Fm Palmar Fm
Sanalejos Vichú
(a) Member Member
(a′)
H2 H9 H11 V33 V90 V95 V97
Betijoque ( –70° 43' 53'',9° 22' 32''; 574 m) 0
Two-way traveltime (s)

1
A
Be
tijo
qu 2
Isno e. F
Palmar F. tú F
.
3

A'
0 5 10 15 20 km
0 20 40 km
Altamira de Barinitas Quebrada Barinas
Cáceres Seca
(b) Pagüey Fm Parángula Fm Río Yuca Fm (b′)
1.5
N-3 1 G2 1 Y5 Y4 3 Y2 Y8
Barinitas (–70°24'51'', 8°45'47''; 1397 m) PA PAN- TP TPGTPMTPMTPMY TPM TP
M
MY
7
B TP
0
Depth/elevation (km)

y. F.
Pagüe –1.5
Parángula .F
Holocene-Pleistocene
–3.0 Plio-Mio-Oligocene
Río Yuca F.
Eocene-Paleocene

B' –4.5 Cretaceous-Jurassic (undiff)


SIR* (Lower to Upper Paleozoic)
* Silicic Intrusive Rock (SIR)
–6.0
0 5 10 15 20 25 30 35 40 45 km
0 20 40 km

Fig. 3. (a) Line drawing of seismic section of the southern part of the Maracaibo basin and the northern flank of the Venezuelan
Andes, showing approximate sample locations along the Rıo Hoyos and Rıo Vich u sections. The seismic profile was provided by Pet-
roleos de Venezuela, SA. (b) Line drawing of seismic section of the Barinas basin showing sample locations along the Parangula River
section near Barinitas on the southern flank of the Venezuelan Andes. The seismic profile was modified from Henriques-Casas (2004).
Vertical scale is in TWT (s) for profile aa’ and in km for bb’ (0 km corresponds to sea level).

© 2015 The Authors


4 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

the Barinas basin was connected to the Maracaibo basin deposits (CH1-CH73, Fig. 5), while two samples of
(Zambrano et al., 1972; Roure et al., 1997; James, 2000; organic material were collected for Rock-Eval analyses
Guzman & Fisher, 2006). Middle Eocene marine clastic (CH51(1), CH52(2); Fig. 5). For AFT analysis, samples
sediments of the Gobernador and Pag€ uey formations were collected from the Eocene Pag€uey Formation (PAN-
reach a thickness of up to 1 km in the Barinas basin. In 3), the Parangula Formation (PAN-1, TPG-1, TPG-2)
contrast with most parts of the Maracaibo basin, the cen- and the Rıo Yuca Formation (TPMY-2, TPMY-3,
tral part of the Barinas basin was affected by subsidence TPMY-4, TPMY-5, TPMY-7 and TPMY-8) along the
during the Oligocene, leading to deposition of the Car- Parangula River section (Fig. 5).
bonera Formation. This formation is not preserved in
the proximal basin, where units are generally tilted to the
south and exposed along the southern flank of METHODS
the Venezuelan Andes today (Fig. 3b). Here, the Eocene
Sediment petrography
Pag€uey Formation is unconformably overlain by the Mio-
cene fluvial Parangula Formation, which in turn is Stratigraphic logs and descriptions of conglomerate pet-
(unconformably?) overlain by the late Miocene to rography are based on our field observations. In the Mara-
Pliocene fluvial Rıo Yuca Formation. In contrast with the caibo basin section, we performed clast counts over a
relatively deep northern pro-wedge foreland basin, the 3600-cm2 surface area directly on the conglomerate out-
southern retro-wedge foreland basin reaches a maximum crops. Only pebbles were counted that had a minimum
thickness of only 2–3 km (De Toni & Kellogg, 1993; size of 3 cm along their long axis, distinguishing granite,
Chacın et al., 2005; Jacome & Chacın, 2008). gneiss, amphibolite quartzite, sandstone and mudstone
clasts. Percentages were calculated from the total clast
counts (about 50 counts per location), normalized to 100
SAMPLE COLLECTION percent. No pebble counts were performed in the Barinas
basin section because of the lack of conglomerates.
Maracaibo basin
Thin sections of sandstone samples were point-counted
On the northern flank of the Venezuelan Andes, two com- at 10–1009 magnification using a BX41 Olympus polariz-
plementary sections were mapped and sampled along the ing microscope; over 300 points were counted per thin
Rıo Hoyos and Rıo Vich u, north of the city of Betijoque, section. We acknowledge that grain-size may bias the
in the state of Trujillo (Figs 2 and 3a). The combined sec- results (Garzanti et al., 2007) but because the samples
tions record about 2700 m of continuous stratigraphy, showed at least moderate sorting, this bias is considered
entirely in the Pliocene Betijoque Formation (Fig. 4). small. Therefore, we distinguished mono- and polycrys-
Seven sandstone samples were collected along the Rıo talline (>3 subgrains) quartz, feldspar (K-feldspar and
Hoyos and Rıo Vich u sections for petrographic thin-sec- plagioclase), mica (muscovite, biotite and chlorite), sedi-
tion analysis (RH2, RH3, RH9, RH17, RV2, RV3 and mentary lithic grains (chert, sandstone, siltstone, mud-
RV6). In addition, 52 palynological samples were col- stone and carbonate clasts), metamorphic lithic grains
lected from mudrock deposits along the Rıo Hoyos sec- (phyllite, mica schist, gneiss), igneous lithic grains (gran-
tion (CH74 to CH 125; Fig. 4), and two samples of ite), other grains, calcite cement, iron oxide cement,
organic material were collected for Rock-Eval analyses matrix and porosity (primary and secondary). The values
(RV1, RV3; Fig. 4). A total of seven samples, ~5–15 kg of mono- and polycrystalline quartz were combined to
each, were collected from medium to coarse-grained calculate total quartz for representation in a Quartz-Feld-
sandstones of the Betijoque Formation for detrital apatite spar-Lithic grains (QFL) diagram (Folk, 1965). The feld-
fission-track (AFT) analysis. Three samples (H2, H9 and spar component is calculated by combining K-feldspar
H11) were collected along the Rıo Hoyos section, and four and plagioclase. Lithic grains were combined for calculat-
samples (V33, V90, V95 and V97) were collected along ing the rock-fragment component.
the Rıo Vichu section (Fig. 4).
Palynology
Barinas basin
The palynological analysis was aimed at better constrain-
The entire section of the Upper Miocene-Pliocene Paran- ing the depositional age of the Betijoque and Rıo Yuca
gula and Rıo Yuca formations was studied along the formations. For this reason, particular attention was paid
Parangula River to the west and south of the town of to the presence of biostratigraphic markers as defined for
Barinitas, on the southern flank of the Venezuelan Andes the Caribbean (Germeraad et al., 1968; Lorente, 1986;
(Figs 2 and 3b). The total thickness of the Neogene sec- Muller et al., 1987), the Andes (Hooghiemstra, 1984;
tion we studied is about 2700 m. Eleven sandstone sam- Torres-Torres, 2006) and the Llanos basin (Jaramillo
ples were collected for petrographic thin section analysis et al., 2011). The most productive samples from the Beti-
(PAS2, PAS6, PAS10, PAS13, PAS20, PAS24, PAS28, joque Formation were collected from a one meter thick
PAS29, PAS35, PAS41 and PAS42, Figs 5 and 9). A total series of grey laminated clays (at 1620 m in our section)
of 73 palynological samples were collected from mudrock that represent a channel infill in a sequence of otherwise

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 5

M.A. Bermudez et al.

during Atlantic rifting (Case et al., 1990; Pindell & deposition of the shallow marine clastic La Rosa/Palmar
Kennan, 2009). These grabens were filled with clastic sed- Formation followed by the Upper Miocene fluvial Isnot u
imentary rocks of the La Quinta Formation. Continuous Formation and the Pliocene alluvial fan deposits of the
rifting led to development of a passive margin during the Betijoque Formation. The Betijoque Formation can be
Early Cretaceous, in which the La Luna Formation, a further divided into the Vich u and Sanalejos members
calcareous shale unit that is the source rock of 98% of all (Guerrero et al., 2009). Proximal parts of the Palmar,
petroleum reserves in the Maracaibo basin (Escalona & Isnot
u and Betijoque formations have been tilted to the
Mann, 2006), was deposited. The La Luna Formation is north and exposed along the northern flank of the
overlain by the Upper Cretaceous Colon Formation and Venezuelan Andes since the late Pliocene-Pleistocene
the Paleocene Guasare Formation. The tectonic setting of (Figs 2 and 3a). In total, the pro-wedge foreland basin of
the Maracaibo basin changed from a passive margin to a the Venezuelan Andes reaches a maximum thickness of
foreland basin in the late Paleocene, due to the onset of about 6 km (De Toni & Kellogg, 1993).
oblique convergence between the Caribbean and South Miocene-Pliocene flexural subsidence was caused by
American plates (Lugo & Mann, 1995). Clastic sediments, crustal loading of the rising Venezuelan Andes, which
derived from the uplifted and eroded northern Colombian resulted from oblique collision of the Maracaibo block
Andes, were transported into the Maracaibo basin, result- with the South American plate (De Toni & Kellogg,
ing in deposition of the Eocene shallow-marine/deltaic 1993). This change in regional plate kinematics may have
Misoa Formation and migration of the basin depocenter been driven by collision of the Panama arc with the north-
to the west. The mainly quartz-arenitic Misoa Formation ern Andes in Colombia (Kohn et al., 1984; Colletta et al.,
is one of the two main reservoir rocks of the Maracaibo 1997), resulting in clockwise rotation of the Maracaibo
basin today (Escalona & Mann, 2006). The reservoir is block and the Santa Marta massif (Fig. 1; Montes et al.,
sealed by low-permeability sediments on top of a regional 2005, 2010).
Late-Eocene/Oligocene unconformity, which indicates The Barinas basin is limited to the southeast by the
uplift and partial erosion of the basin deposits during the Guyana shield, to the west by the Eastern Cordillera of
Oligocene. Subsidence recommenced in the southern the Colombian Andes, and to the northeast by the El Ba ul
Maracaibo basin in the Early-Middle Miocene, with Arch (Fig. 1). During the Late Cretaceous and Eocene,

Cerro Alto
NW Sabana de Betijoque Isnotú SE
Mendoza Betijoque Fm Isnotú Fm Palmar Fm
Sanalejos Vichú
(a) Member Member
(a′)
H2 H9 H11 V33 V90 V95 V97
Betijoque ( –70° 43' 53'',9° 22' 32''; 574 m) 0
Two-way traveltime (s)

1
A
Be
tijo
qu 2
Isno e. F
Palmar F. tú F
.
3

A'
0 5 10 15 20 km
0 20 40 km
Altamira de Barinitas Quebrada Barinas
Cáceres Seca
(b) Pagüey Fm Parángula Fm Río Yuca Fm (b′)
1.5
N-3 1 G2 1 Y5 Y4 3 Y2 Y8
Barinitas (–70°24'51'', 8°45'47''; 1397 m) PA PAN- TP TPGTPMTPMTPMY TPM TP
M
MY
7
B TP
0
Depth/elevation (km)

y. F.
Pagüe –1.5
Parángula .F
Holocene-Pleistocene
–3.0 Plio-Mio-Oligocene
Río Yuca F.
Eocene-Paleocene

B' –4.5 Cretaceous-Jurassic (undiff)


SIR* (Lower to Upper Paleozoic)
* Silicic Intrusive Rock (SIR)
–6.0
0 5 10 15 20 25 30 35 40 45 km
0 20 40 km

Fig. 3. (a) Line drawing of seismic section of the southern part of the Maracaibo basin and the northern flank of the Venezuelan
Andes, showing approximate sample locations along the Rıo Hoyos and Rıo Vich u sections. The seismic profile was provided by Pet-
roleos de Venezuela, SA. (b) Line drawing of seismic section of the Barinas basin showing sample locations along the Parangula River
section near Barinitas on the southern flank of the Venezuelan Andes. The seismic profile was modified from Henriques-Casas (2004).
Vertical scale is in TWT (s) for profile aa’ and in km for bb’ (0 km corresponds to sea level).

© 2015 The Authors


4 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

Parangula River section Pebble count Pollen AFT


analysis samples samples
Stratigraphic Top
thickness (m) of section
0 TPMY7

TPMY8

PLIOCENE
Bridge
PAS20
PAS22 CH18–19
PAS21
PAS23 CH7–17

500 CH54–65
UPPER

R I O Y U C A Fm

TPMY2
PAS24
PAS25
PAS26

TPMY3
PAS27

PAS28 CH3–4

Bridge PAS30 PAS29 CH1–2


1000
PAS31
PAS32
PLIOCENE

PAS33
PAS34 PAS35
PAS37
PAS36

CH47–54 Rock-Eval sample (CH52)


PAS38 TPMY4

PAS39 CH20–46
PAS40

1500
LOWER

CH66
PAS42

PAS13 PAS12 TPMY5


PAS41
PAS11
-

P A R A N G U L A Fm
UPPER MIOCENE

2000

PAS10
PAS9

PAS4 PAS5
PAS8
PAS7

PAS3
CH67–71
PAS6 TPG1
PAS2
2500
TPG2
PAS1
Bridge

PAN 1 PAN 1
CH72–73
Eocene Pagüey
Fm. PAN 3 PAN 3

Fig. 5. Generalized stratigraphic section of the Parangula and Rıo Yuca formations from the Barinas basin on the southern flank of
the Venezuelan Andes. Stratigraphic positions of samples collected for different analysis (sandstone petrography, palinology, apatite
fission-track analysis) are indicated. Samples for Rock-Eval analysis are highlighted in dark-grey.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 7

M.A. Bermudez et al.

braided fluvial deposits. This bed was sampled at every 10 fission-track grain-age distributions were analyzed for
centimeters and all samples were processed. The samples obtaining central ages (Galbraith & Laslett, 1993) and
from the Rıo Yuca Formation were collected from very decomposed into major grain-age components or peak
organic-rich clay beds that occur in the middle part of the ages, using binomial peak-fitting as described in Stewart
formation. The palynological slides were rich in fine, & Brandon (2004). Although there are several methods
black and dark-brown coloured organic matter; neverthe- available to decompose mixtures ages (Bermúdez et al.,
less, pollen densities were very low. Of 31 samples from 2015), we found no significant differences between these
the Rıo Yuca Formation that were chosen for processing, methods. The obtained AFT peak ages can be related to
25 samples contained pollen but only six slides produced certain source areas in the Venezuelan Andes by compar-
sufficient pollen for analysis (above 50 grains per slide). ison with published bedrock AFT ages (Kohn et al.,
The palynological assemblage was dominated by 1984; Berm udez et al., 2010, 2011).
reworked pollen species. For this reason the palynological
results are represented in table format, and not in the
usual graphical manner. The positions of all samples are
Rock-Eval analysis
indicated in the stratigraphic columns in Figs 4 and 5. Rock-Eval analyses of organic matter were performed by
Palynological processing followed standard procedures F. Baudin at Universite Pierre and Marie Curie in Paris,
at the department of Paleoecology and Landscape Ecol- France, using an Oil Show Analyzer and following the
ogy at the University of Amsterdam. For each sample 1 procedure described by Espitalie et al. (1986). The
cm3 of material was taken. Organic-rich clays and silts organic carbon composition of a sedimentary rock can be
were treated with a 10% sodium pyrophosphate used as a proxy for the maximum temperature experi-
(Na4P2O7.10H2O) solution, acetolysis, and sieved over a enced during burial. This method requires heating of
212 lm mesh size. At the end of the process bromoform about 100 mg of sample material to volatize the organic
with a density of 2.0 g cm 3 was used to separate the carbon contained in the sample and to expel petroleum.
organic and inorganic fractions, and the organic residue The amount of petroleum expelled increases with an
was mounted in glycerin and sealed with paraffin. increasing hydrogen to carbon ratio (H/C). The results
The sporomorph content of each slide was counted are plotted in a Rock-Eval Tmax (°C) vs. Hydrogen Index
applying morphogeneric names, but in some cases affini- (HI) graph; the Hydrogen Index is defined as HI = (S2/
ties with modern taxa was indicated. All taxonomic refer- TOC)*100 (Espitalie et al., 1977), where TOC is the per-
ences and age ranges (based on Jaramillo et al., 2011) are centage of Total Organic Carbon, and S2 is the residual
listed in the supplementary data to this manuscript. For petroleum potential produced at Tmax, which is depen-
this work we consulted, among others, Gonzalez-Guzman dent on the type of kerogen, this could be to explain the
(1967), Germeraad et al. (1968), Hooghiemstra (1984), different paths obtained when plotting Tmax vs. HI.
Muller et al. (1987), Lorente (1986), Jaramillo & Rueda Samples that plot within the immature field on the left of
(2008) and Jaramillo et al. (2011). the graph indicate that maximum burial temperature con-
ditions remained below the oil window. This coincides
with part of the AFT partial annealing zone (~60–120°C).
Detrital thermochronology
Detrital apatites were extracted from sandstone samples
using sieving, heavy-liquid and magnetic separation tech- RESULTS
niques. During mounting, 500–800 grains were randomly
distributed on the mounts to avoid sampling bias during Sedimentology and sedimentary petrography
counting, given the variety of grains in a detrital sample
Maracaibo basin
with respect to their cooling age and uranium content.
Apatite aliquots were mounted in epoxy resin and polished Because of limited accessibility and poor outcrop condi-
to expose internal grain surfaces. The apatite grain mounts tions, fieldwork focused on the upper part of the Vich u
were etched for 20 s in 5.5 M HNO3 at 21 °C. The samples Member and most of the Sanalejos Member of the Betijo-
were irradiated together with Fish Canyon Tuff and Dur- que Formation (Fig. 4). For most of the section the strata
ango apatite age standards and IRMM-540 uranium glass strike and dip consistently at about 330/35 NW. However,
dosimeters at the well-thermalized Orphee reactor, Centre at the top of the section the strata rapidly become sub-hori-
d’Etudes Nucleaires de Saclay, Gif-sur-Yvette, France. zontal (Fig. 3a). While the upper part of the Vich
u Member
Randomly selected grains were counted dry at 12509 contains several meter-thick alternating fine sandstone and
magnification using an Olympus BH-2 microscope. Our mudrock deposits, the Sanalejos Member is marked in gen-
goal was to date about 100 grains per sample, to achieve eral by 1 to 2 m thick, mainly matrix supported polymict
the required level of statistical adequacy for provenance conglomerate beds, alternating with sandstone and
studies (Vermeesch, 2004). However, it was not possible mudrock layers (Fig. 6). Conglomerate clasts of the Sanale-
to find this number of countable grains in all samples jos Member include gneiss, granite, amphibolite, sandstone,
because of inclusions, dislocations and poor apatite mudrock and vein-quartz pebbles, with granite decreasing
quality caused by intense tropical weathering. Observed upsection (see pebble count diagrams in Fig. 4).

© 2015 The Authors


8 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists

M.A. Bermudez et al.

Rio Hoyos - Rio Vichu section Pebble count analysis


Stratigraphic Top
thickness (m) of section
0 Pebble Pollen AFT
count samples samples
analysis

500

Pebble count location RH18


Sanalejos Member

13%
36%

1000

RH21 CH117–125 32%


H2
Formation

0% 8%
11%
PLIOCENE

RV9 Pebble count location RH14


BETIJOQUE

22% 0%
RH20 CH115–116
RH19 CH114 H9 40%
RH18 CH105–113
1500
RH17 CH97–104
RH16 CH95–96
RH15 CH93–94
RV8 20%
RH14 CH91–92 4% 14%
RV7
RH13 CH90
RH12 CH89
RH11 RV6 CH87–88

RH10 CH86 H11


Pebble count location RH11
RH9 CH82–85 0%
RH8 CH81 V33 16%
RH7 CH80
RH6 CH79 10%
2000 RH5 CH78
RH4 CH76–77
RH3 CH75 54%
RH2
RH1 CH74 8%
RH0 RV5
12%
V90
RV4

RV3 Rock-Eval sample (RV3)


Pebble count location RH9 Granite
Vichu Member

12% 0% Gneiss
V95 Amphibolite
Quarzite
2500
25% 51% Sandstone
RV2 Mudstone
6%
RV1 Rock-Eval sample (RV1) 6%
RV0
V97

Legend Coarse sandstone Mudstone

Conglomerate Fine sandstone No outcrop Fossils

Fig. 4. Generalized stratigraphic section of the Betijoque Formation from the Maracaibo basin on the northern flank of the Venezue-
lan Andes; Stratigraphic positions of samples collected for different analysis (sandstone petrography, palynology, apatite fission-track
analysis) are indicated. Samples for Rock-Eval analysis are highlighted in dark-gray. Pebble counts are presented from four locations.

© 2015 The Authors


6 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists

M.A. Bermudez et al.

coarse-grained sandstone of the Rıo Yuca Formation is at a paramo vegetation, which is characteristic of altitudes
observed in the field (Fig. 8c,d). From about the middle between 3500 and 4000 m (e.g. Hooghiemstra, 1984;
of the Rıo Yuca Formation upsection, matrix-supported van’t Veer & Hooghiemstra, 2000; Rull, 2006).
polymict conglomerate layers appear (Fig. 8e,f). These The palynological assemblage further contains fern
layers are more frequent towards the top of the section. spores such as Kuylisporites waterbolkii (Cyathea horrida)
Similar to the Betijoque Formation in the Maracaibo and Nijssenosporites sp. (aff. Pityrogramma) and pollen
basin, the conglomerates include gneiss, granite, sand- such as Clavainaperturites microclavatus (Hedyosmum),
stone, and vein-quartz clasts (Fig. 8a,b). aff. Symplocos (Apocynaceae), aff. Draba and Podocar-
Thin-section analysis and point counting show that the pidites sp. All these taxa are typical for the Andean forest
sandstones of the Parangula and Rıo Yuca formations are at 2300–3000 m elevation (e.g. see Hooghiemstra, 1984;
mostly relatively quartz-rich lithic sandstones (Fig. 9a). van’t Veer & Hooghiemstra, 2000; Rull, 2006). In con-
The two samples from the Parangula Formation (PAS6-2 trast, humid tropical lowland taxa were almost absent
and PAS10) have much higher quartz concentrations than from the palynological assemblage.
the samples from the Rıo Yuca Formation and contain The low pollen numbers may be related to the deposi-
only sedimentary lithoclasts (Table 1b). The observed tional environment in which the sediments were formed.
lithic grains in the Rıo Yuca Formation include granitic Palynological sampling in braided-fluvial to alluvial-fan
and gneissic basement rock fragments, as well as sedimen- sediments is challenging as silt and clay, the fractions in
tary lithic grains (Fig. 9b–d). which pollen are generally found, are rare. Another factor
hampering the palynological research in these kinds of
depositional environments is the damage caused to the
Palynology pollen wall by repetitive oxidizing conditions.
Maracaibo basin
Barinas basin
Only two samples (Table S1) from the Betijoque Forma-
tion produced acceptable quantities of pollen; palynologi- The pollen assemblage contained biostratigraphically sig-
cal results should thus be regarded with some caution. nificant taxa such as Cyatheacidites annulatus, Grimsdalea
Samples CH100 and CH105 from the middle part of magnaclavata, Crassoretitriletes vanraadhooveni, Psi-
the section (locations RH18 and RH19 in Fig. 4a, lastephanocolpites evansii, Striasyncolpites zwaardii, Echitri-
Table 2a) contain Cyatheacidites annulatus and Lana- colporites spinosus and Fenestrites spinosus. Based on the
giopollis crassa. The former has a First Appearance Date presence of C. annulatus, G. magnaclavata and C. van-
of 7.15 Ma and the latter has a reliable Last Appearance raadshooveni, the age is estimated as ranging from 7.15
Date of 4.77 Ma, suggesting a late Miocene to Pliocene Ma to 3.4 Ma (late Miocene-early late Pliocene). The
age (Jaramillo et al., 2011). Key markers such as Grims- absence of E. mcneilly (<1.56 Ma), Myrica (first occur-
dalea magnaclavata and Crassoretitriletes vanraadshooveni rence late Pliocene) and Alnus (first occurrence Middle
(both >3.4 Ma), E. mcneilly (<1.56 Ma), Myrica (late Pleistocene) further supports this age estimate.
Pliocene) and Alnus (Middle Pleistocene) are absent We also found that a significant proportion of the pol-
(Fig. 10). In addition, taxa with affinity (aff.) Valeriana, len assemblage was composed of Lanagiopollis crassa and
Ericipites sp., Foveotriletes cf. ornatus (aff. Huperzia) point occurrences of dinoflagellates and inner (organic) linings

Quartz Sedimentary
(a) 0
100
(b) 0
100
10 10
90 90
20 20 RH17
80 80
30 30 Recycling of
70 70
RH9 RV6
40 40 sedimentary
RV2 60 60
50 RV6 50 cover
RV3 50 50
RH2
60 60 RH9
RH17 40 40
70 RH2
RH3 30 70
30
RH3
80 80
20 20
90 RV2
10 90
10
RV3
100 100 Crystalline basement
0 0
Feldspar 0 10 20 30 40 50 60 70 80 90 100 Lithic grains Plutonic 0 10 20 30 40 50 60 70 80 90 100 Metamorphic

Fig. 7. Triangular plots showing normalized point-count data of the Betijoque Formation. (a) QFL plot; Q is total quartzose grains;
F is total feldspar grains; L is total unstable lithic fragments; (b) Lithic grains plot showing plutonic lithic fragments, metamorphic
lithic fragments and sedimentary lithic fragments.

© 2015 The Authors


10 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

Fig. 8. Outcrop photos form the Barinas


basin section. (a) Location PAS8 in the
Parangula Formation, showing this over-
bank mudstone and channel sandstones.
(b) Location PAS6 in the Parangula For-
mation. Black arrows mark angular
unconformity with overlying Quaternary
conglomerates and soil layers. (c) Massive
layers medium grained sandstone of the
Rıo Yuca Formation at location PAS 24.
(d) Massive layers of channel sandstones
in the Rıo Yuca Formation at location
PAS 28. Note the erosive contact shown
with black arrow. (e) Conglomerate layer
with sand lenses in the upper part of the
Rıo Yuca Formation, location PAS29. (f)
Conglomeratic deposits in grey, quartz-
rich sandstone at location PAS 20 of the
upper Rıo Yuca Formation.

of foraminifers. Dinoflagellate cysts are few and poorly addition, the presence of diverse taxa typical for the
preserved in the samples from the Rıo Yuca Formation in Andean forest such as Podocarpidites (Podocarpus) and
the Barinas basin section. Polysphaeridium spp. is the most Clavainaperturites microclavatus (Hedyosmum), and some
frequent, followed by Spiniferites spp. Rare or single paramo taxa like Foveotriletes cf. ornatus type (aff.
occurrences include Areoligera sp., Homotryblium sp. and Huperzia) and Nijssenosporites sp. (aff. Pityrogramma)
a questionable specimen of Diphyes colligerum. The indicate the existence of a montane vegetation. The
Polysphaeridium specimens bear numerous slender pro- assemblage is low in spores and thus ferns must not have
cesses, which is more typical for the ‘older’ types, particu- formed a significant part in the paleovegetation.
larly for the Eocene-Oligocene types. The combined Finally, the abundance of clay beds rich in fossil leaves
presence of Homotryblium, Areoligera and the question- offers good perspectives for a reconstruction of the local
able specimen of Diphyes colligerum, would indicate an flora at the time of deposition. Fragments of palm leaves
Eocene age (Powell, 1992; Brinkhuis & Biffi, 1993; Bujak and Melastomataceae suggest that the local vegetation was
& Mudge, 1994; Zevenboom, 1995), but the poor preser- of humid, tropical character. Further investigation and
vation and the scattered occurrences suggest reworking. sampling could yield valuable new data on the late
The presence of pollen taxa such as Spinozonocolpites Miocene-early late Pliocene climate in this region.
sp., Spirosyncolpites spiralis, Annutriporites iversenii,
Retibrevitricolpites sp. and Cicatricosisporites dorongensis,
among others, suggests reworking from pre-Miocene Detrital thermochronology
units and in particular of the Eocene-Oligocene forma-
Maracaibo basin
tions. In addition, the occurrence of elaterate pollen also
indicates reworking from Cretaceous sediments. The A total of 444 individual apatite grains from the seven
palynological data are listed in Table S1; some of the most sandstone samples collected in the field were dated with
characteristic taxa are illustrated in Fig. 11. the fission-track method. All samples fail the v2 test, indi-
Other components of the palynological assemblage, cating that they are characterized by multiple age compo-
such as Bombacacidites (Bombacaceae), Retitrescolpites? nents or peaks. Consequently, apatite grains are
irregularis (Amanoa) and Mauritiidites (Mauritia palm), interpreted as being derived from multiple sources. The
point at the existence of tropical lowland vegetation. In age range of dated grains and binomial fitted peaks of all

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 11

M.A. Bermudez et al.

(a) (b)

(c) (d)

Fig. 9. Triangular plots showing normalized point count data of the Parangula and Rıo Yuca formations. (a) QFL plot; Q is total
quartzose grains; F is total feldspar grains; L is total unstable lithic fragment; (b) Lithic grains plot showing plutonic lithic fragments,
metamorphic lithic fragments and sedimentary lithic fragments. (c) Example of a granite lithic grain (PAS20). (d) Example of a mica
schist lithic grain (PAS 20).

seven samples are reported in Table 2. Probability density are shown in Fig. 13. Sample PAN-3 from the Eocene
plots of all seven samples are shown in Fig. 12. All samples Pag€uey Formation was of very poor quality and only five
from the Betijoque Formation except sample V95 contain grains could be dated from this sample, which give a late
late Miocene to Early Pliocene age components, with a Miocene-Early Pliocene central age (6.4  1.8 Ma).
peak age ranging between 4 and 8 Ma. All samples further This result is similar to the 5.0  1.0 Ma central age
contain apatites with mid-late Miocene cooling ages (age determined from 25 grains of sample PAN-1, from the
components of 10–18 Ma), which are the youngest ages in very base of the Parangula Formation. Samples TPG-1
sample V95. In addition, Eocene to early Miocene age com- and TPG-2, stratigraphically higher up in the Parangula
ponents can be discriminated in several samples, while Formation, have central ages younger than their suppos-
sample H9 also shows a minor Jurassic age component edly early Miocene depositional age. Sample TPG-1
(Table 2). Only few track lengths could be measured in from the Parangula Formation and all samples from the
these samples; where more than 10 lengths could be mea- Rıo Yuca Formation show late Miocene age components
sured these vary between 12.2 and 13.1 lm (Table 2). of about 11-9 Ma, and most samples also have an Early
to Middle Miocene peak of about 22 to 15 Ma. Oligo-
cene and older age peaks are only present in samples
Barinas basin
TPG-1, TPMY-5, TPMY-4 and TPMY-7 (Table 2).
Well over 900 individual apatite grains were dated for Sample TPMY-8 from near the top of the studied sec-
the samples of the Barinas basin. The age range of the tion has young peak age of 4.6  0.9, comparable to
dated grains and binomial fitted peaks are reported in detrital AFT ages in samples from the Maracaibo basin
Table 2. With the exception of samples PAN-1, PAN-3 (Table 2) and in-situ bedrock AFT ages in the present-
and TPG-2, which yielded very low numbers of count- day Santo Domingo River catchment (Berm udez et al.,
able grains, all samples failed the v2 test and are thus 2010, 2013). Reasonable numbers of track lengths could
characterized by multiple grain-age components, indicat- be measured in six samples and, like for the Maracaibo
ing that grains were derived from several different basin samples, mean track lengths vary between 12.2 and
sources. Probability density plots with best-fit peak ages 13.3 lm (Table 2).

© 2015 The Authors


12 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Table 2. Detrital apatite fission-track results of the Maracaibo and Barinas basins, Venezuela

Stratigraphic Age range Central age 


Sample depth (m) n (Ma) P1 (Ma) P2 (Ma) P3 (Ma) P4 (Ma) P5 (Ma) 1r (Ma) MTL  SE (lm) SD (lm) Dpar  SD (lm)

Maracaibo

© 2015 The Authors


H02 1075 102 2.2–39.7 6.2  0.7 17.2  3.0 – – – 7.8  1.0 12.90  0.31 (16) 1.23 1.79  0.41
85.3% 14.7%
H09 1425 79 2.6–161.6 4.5  2.7 10.0  2.7 23.7  4.5 38.7  6.9 163.6  125 15.3  3.1
23.4% 41.4% 18.9% 15.1% 1.3%
H11 1810 100 3.4–67.0 8.2  1.9 15.4  2.2 33.8  7.4 – – 15.4  1.9 12.19  0.34 (27) 1.76 1.70  0.22
27.9% 59.2% 12.9%
V33 1900 49 3.9–49.3 5.1  1.2 11.1  1.1 41.8  36.5 – – 10.4  1.3 13.08  0.25 (26) 1.30 2.95  0.45
16.2% 81.0% 2.7%
V90 2160 40 2.5–67.9 4.4  1.0 18.3  4.6 40.5  28.8 – – 16.6  4.3 9.16 (01) – 1.34
27.9% 54.5% 17.6%
V95 2425 50 3.5–61.3 12.0  1.9 28.1  19.8 – – – 13.8  2.0 9.79  0.002 (02) 0.003 1.35  0.20
85.7% 14.3%
V97 2700 24 4.4–29.3 7.0  1.6 18.7  25.4 – – – 7.3  1.4 12.70  1.56 (03) 2.71 2.85  0.39
92.9% 7.1%
Barinas
TPMY7 0 40 4.8–100.0 10.3  1.5 96.9  74.3 – – – 12.2  2.8 12.88  0.25 (22) 1.17 1.99  0.25
97.4% 2.6%
TPMY8 60 100 2.2–34.4 4.6  0.9 9.1  2.1 19.9  22.6 – – 7.3  0.8 12.96  0.38 (16) 1.53 1.52  0.15
42.0% 54.8% 3.2%
TPMY2 650 100 3.8–35.0 11.6  2.2 78.2% 16.1  7.9 – – – 12.6  1.4 – – –
21.8%
TPMY3 860 103 4.2–76.4 11.1  1.4 22.1  3.3 – – – 15.6  1.6 13.31  0.41 (12) 1.43 1.66  0.17
56.2% 43.8%
TPMY4 1260 364 2.9–120.8 9.5  1.2 15.9  1.8 32.7  5.2 – – 15.6  1.1 12.52  0.15 (62) 1.22 2.84  0.46
33.0% 52.2% 14.8%
TPMY5 1700 120 4.9–29.6 9.4  2.3 13.3  2.4 24.8  6.8 – – 12.2  1.3 12.29  0.15 (48) 1.05 1.78  0.20
38.0% 58.3% 3.7%
TPG1 2450 100 3.5–54.6 9.6  0.9 18.7  4.0 53.0  42.8 – – 11.4  1.1 12.16  0.18 (60) 1.38 1.68  0.20
79.5% 19.6% 0.9%
TPG2 2550 10 7.5–41.6 – – – – – 14.2  3.3 – – –
PAN 1 2650 25 3.4–12.1 – – – – – 5.0  1.0 – – –
PAN 3 Pag€
uey Fm. 5 1.9–9.5 – – – – – 6.4  1.8 – – –

Note: n = total number of grains counted; binomial fitted peak ages are given 2 SE. No peaks were fitted for sample TPG2 because of the low grain count. Also given is the percentage of grains in a specific peak. All sam-

Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
ples were counted at 12509 dry (1009, 1.25 tube factor, 10 oculars) by M. A. Bermudez using a zeta (IRMM 540) of 288.66  5.32 (1 SE).
Surface uplift and exhumation in basins of the VA

13

M.A. Bermudez et al.

Fig. 10. Characteristic pollen and spores from the Betijoque Formation (Maracaibo Basin) (Table S2). The sample code is indicated
after the taxonomic name: 1. Podocarpidites sp. (Bet 105.1), 2. Monoporopollenites annulatus (Bet 105.2), 3. Clavainaperturites micro-
clavatus (Bet 105.2), 4. aff. Valeriana (Bet 105.2), 5. aff. Symplocos (Bet 105.2), 6. Spirosyncolpites spiralis (Bet 105.2), 7. Echitricolporites
spinosus, 8. Ericipites sp., 9. Verrucatotriletes cf. bullatus (Bet 105), 10a & b. Retitriletes sp. (105.2), 11. Cyatheacidites annulatus (Bet
105.1), 12. Cingulasporites laevigatus (Bet 105.1), 13a & b. Foveotriletes cf. ornatus (Bet 105.1), 14. indeterminate trilete spore (Bet
105.1), 15. Psilatriletes lobatus (Bet 105.2), 16. Nijssenosporites fossulatus (Bet 105.1), 17. Nijssenosporites sp., fragment of characteristic
ridges (Bet 105.1) and 18. Matonisporites mulleri (Bet 105.2).

© 2015 The Authors


14 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

Fig. 11. Characteristic pollen, spores and dinoflagellates in the Rıo Yuca Formation (Barinas Basin) (Table S1). The sample code is
indicated after the taxonomic name:1. Bombacacidites brevis (CH 40.2), 2. Psilaperiporites robustus (CH 26), 3. Glencopollis curvimuratus
(CH 40.2), 4a. Crototricolpites sp., 4b. Detail of sculpture (CH 40.2), 5. Retibrevitricolpites sp. (CH 40.2), 6. Cyclusphaera scabrata (CH
40.2), 7. Lanagiopollis crassa (CH 40.2), 8. Lanagiopollis crassa (CH 40.2), 9. Nijssenosporites sp. (CH 40.2), 10. Cicatricosisporites doron-
gensis (CH 40.2), 11. Spiniferites sp. (CH 38), 12. Lingulodinium machaerophorum? (CH 40.2), 13. Spiniferites sp. (CH 50.1), 14. indeter-
minate dinoflagellate (CH 40.2) and 15. foraminifer lining (CH 44).

percentages of total organic carbon (TOC) levels, in the


Rock-Eval analysis range of coal (Baudin et al., 2007). Both samples plot
The Rock-Eval analyses of organic matter of samples RV1 below the field of hydrocarbon maturation in the Rock-
and RV3 from the Maracaibo Basin indicate significant Eval Tmax (°C) diagram shown in Fig. 14. The other

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 15

M.A. Bermudez et al.

22

Probability density (%/Δz = 0.1)


20 V33 P1 = 5.1 ± 1.2
18 N = 50 P2 = 11.1 ± 1.1
P3 = 41.8 ± 36.1
16
14 P2
12
10
8
6
4 P1 P3
2
0
1 2 3 5 7 10 20 30 50 100

12 11

Probability density (%/Δz = 0.1)


Probability density (%/Δz = 0.1)

11 H2 P1 = 6.2 ± 0.7 10 V90


10 N = 105 P2 = 17.2 ± 3.0 9 N = 40
9 8 P1 = 4.4 ± 1.0
8 P1 7 P2 = 18.3 ± 4.5
7 6 P3 = 40.5 ± 29.3 P2
6
5
5
4
4 P1
3 3
P2 2 P3
2
1 1
0 0
1 2 3 5 7 10 20 30 50 100 1 2 3 5 7 10 20 30 50 100

8 12
Probability density (%/Δz = 0.1)
Probability density (%/Δz = 0.1)

P1 = 4.5 ± 2.6 H9 11 V95 P1 = 12.0 ± 1.8


7 P2 = 10.0 ± 2.7 N = 79 10 N = 53 P2 = 28.1 ± 19.3
6 P3 = 23.7 ± 4.5 9
P4 = 38.7 ± 6.9 8 P1
5 P5 = 163.6 ± 130.9
7
4 P2 6
5
3 P3
P4 4
2 P1 3
2 P2
1
P5 1
0 0
1 2 3 4 5 10 30 50 100 200 1 2 3 5 7 10 20 30 50 100

10 14
Fig. 12. Detrital AFT results of samples
Probability density (%/Δz = 0.1)
Probability density (%/Δz = 0.1)

9 H11 13 V97 P1 = 7.0 ± 1.6 from the Maracaibo basin on the north-
8 N = 100
12 N = 24 P2 = 18.7 ± 23.9
ern flank of the Venezuelan Andes. The
11
7 10
P2 9 observed grain-age distributions are
6 8
5 P1 = 8.2 ± 1.9 7
P1 shown as histograms and as probability-
4 P2 = 15.4 ± 2.2 6 density functions (continuous black
P3 = 33.8 ± 7.4 5
3 P1
P3
4 lines). Also shown are the binomial best-
2 3
2 P2
fit solutions, with 2–4 peaks. In terms of
1 1
0 0
stratigraphic position the histograms
1 2 3 5 7 10 20 30 50 100 1 2 3 5 7 10 20 30 50 100 should be read down the left hand side
FT grain age (Ma) FT grain age (Ma) and then down the right hand side.

samples collected from the Betijoque Formation did not


contain sufficient organic material for successful analysis
DISCUSSION
(Table 3).
Two samples from the middle of the Rıo Yuca For-
Maracaibo basin
mation in the Barinas basin (Table 3) show TOC levels Sediment provenance
of 0.95 (CH52-1) and 4.74% (CH52-2). Fig. 14 shows
Sandstone petrography and pebble lithologies are good
that both samples plot in the immature field, indicating
indicators of sediment provenance. During the Miocene-
that maximum burial temperatures in the Rıo Yuca
Pliocene, the southern part of the Maracaibo basin, which
Formation were below the temperatures needed for
is the pro-side of the Venezuelan Andes, received detritus
hydrocarbon maturation and therefore below the tem-
derived from the crystalline core of the mountain belt via
perature required for total annealing of fission tracks in
a paleo-drainage system that must have resembled the
apatite.

Fig. 13. Detrital AFT results of samples from the Barinas basin on the southern flank of the Venezuelan Andes. The observed grain-
age distributions are shown as histograms and as probability-density functions (continuous black lines). Also shown are the binomial
best-fit solutions, with 1–3 peaks. In terms of stratigraphic position the histograms should be read down the left hand side and then
down the right hand side.

© 2015 The Authors


16 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

14 16
13 TPMY-7 15 TPMY-5 P1 = 9.4 ± 2.3

Probability density (%/Δz = 0.1)

Probability density (%/Δz = 0.1)


P1 = 10.3 ± 1.5
12 14 P2 = 13.3 ± 2.4
11
N = 40 P2 = 96.6 ± 69.3 13 N = 120
P1 12 P3 = 24.8 ± 6.8
10 11
9 P2
10
8 9
7 8 P1
6 7
5 6
4 5
3 4
3
2 2 P3
1 P2
1
0 0
1 2 3 5 7 10 20 30 50 100 1 2 3 5 7 10 20 30 50 100
16 16
15 P1 = 4.6 ± 0.9 15 TPG-1 P1 = 9.6 ± 0.9

Probability density (%/Δz = 0.1)


Probability density (%/Δz = 0.1)

14 TPMY-8 14
13 N = 100
P2 = 9.1 ± 2.1 13 N = 100 P2 = 18.7 ± 4.0
12 P3 = 19.9 ± 22.6 12 P1 P3 = 53.0 ± 42.8
11 11
10 10
9 P2 9
8 8
7 7
6 6 P2
5 P1 5
4 P3 4
3 3
2 2 P3
1 1
0 0
1 2 3 5 7 10 20 30 50 100 1 2 3 5 7 10 20 30 50 100
14 22
13 P1 = 11.5 ± 2.2 20 TPG-2 Central age =
Probability density (%/Δz = 0.1)
Probability density (%/Δz = 0.1)

TPMY-2
12
11 N = 100 P2 = 16.1 ± 7.9 18 N = 10 14.2 ± 3.3
10 16
9 14
8 12
7 P1
6 10
5 8
4 6
3
4
2
1 P2 2
0 0
1 2 3 5 7 10 20 30 50 100 1 2 3 5 7 10 20 30 50 100
10
P1 = 11.1 ± 1.4 20 PAN-1
9 TPMY-3
Probability density (%/Δz = 0.1)

Central age =
Probability density (%/Δz = 0.1)

P2 = 22.1 ± 3.3 N = 25
8 N = 103
18 5.0 ± 1.0
16
7
14
6 P1
12
5 10
P2
4 8
3 6
2 4
1 2
0 0
1 2 3 5 7 10 20 30 50 100 1 2 3 5 7 10 20 30 50 100
10 42
PAN-3
9 TPMY-4 P1 = 9.5 ± 1.2
Probability density (%/Δz = 0.1)

Probability density (%/Δz = 0.1)

Central age =
N=5
8 N = 364
P2 = 15.9 ± 1.8 35 6.4 ± 1.8
P3 = 32.7 ± 5.2
7
28
6 P2
5 21
4 P1
14
3
P3
2 7
1
0 0
1 2 3 5 7 10 20 30 50 100 1 2 3 5 7 10 20 30 50 100
FT grain age (Ma) FT grain age (Ma)

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 17

M.A. Bermudez et al.

sedimentary lithic grains have three possible sources: (A)


Mesozoic and Paleozoic sedimentary cover of the

Ty

0.5
pe
Venezuelan Andes, (B) Eocene shallow marine deposits,

I
750
such as the Palmar Formation and (C) recycling of poorly
consolidated Middle to late Miocene foreland basin
Hydrogen Index (mg HC g–1 TOC)

450 deposits.

Vitrinite
450
isoreflectance Implications of palynological results
Ty
pe
II The depositional ages of Mio-Pliocene sedimentary rocks
in the Maracaibo basin have hitherto not been well-
300
RV3 constrained. Our palynological analyses provide some age
constraints, suggesting a maximum age of 7 Ma and a
minimum age of 4.77 Ma for the middle of the studied
1.3

150
Betijoque Formation section; a late Miocene-Early Plio-
I
e II cene age appears most likely. Significantly, the palynolog-
Typ
RV1
CH52 (2)
0 CH52 (1) ical spectrum includes taxa that are representative for the
380 400 420 440 460 480 500 520 ‘paramo’, a type of Andean vegetation that is characteris-
Rock Eval Tmax (°C)
tic for altitudes between 3500 and 4000 m (cf. Hooghiem-
Immature Oil window Gas window
stra, 1984; van’t Veer & Hooghiemstra, 2000; Rull, 2006).
Fig. 14. Hydrogen Index (HI) vs. Maximum Temperature Such taxa are known in the Colombian Andes only from
(Tmax) plot for samples RV1 and RV2 from the Betijoque For- the Middle Pleistocene onwards (see Hooghiemstra, 1984;
mation, Maracaibo basin, and for sample CHS2 from the Rıo Torres-Torres, 2006 and references herein). In contrast,
Yuca formation, Barinas basin. None of the samples were suffi- the presence of these taxa in our samples confirms that
ciently heated during burial to allow kerogen maturation. the Venezuelan Andes had already gained a considerable
altitude of at least 3500–4000 m when the Betijoque For-
present-day drainage, as shown by gneiss and granite mation was deposited in the late Miocene-Pliocene. In
lithic grains in the sandstones, and gneiss, granite and addition, the presence of an assemblage of spores and pol-
amphibolite pebbles in the conglomerates (Figs 6a,b). len typical for Andean forest taxa (2300–3000 m altitude;
The variety of granitic pebbles observed in the field Hooghiemstra, 1984; van’t Veer & Hooghiemstra, 2000;
includes micro-granular and epi-granular granite with Rull, 2006) are further evidence that - at the time of depo-
large feldspars and/or hornblende. Further petrographic sition - the vegetation in the source area was distributed
work on these clasts is needed to relate them more closely following an altitudinal zonation.
to Mesozoic, Paleozoic and Proterozoic crystalline rocks
currently exposed in the core of the mountain belt. Beside
Detrital thermochronology
the crystalline fragments, sedimentary lithic grains are
common. The appearance of sedimentary rock fragments The detrital AFT ages in the pro-wedge foreland basin
is an indication of sediment recycling, particularly during (Table 2) correspond closely to the bedrock ages in the oro-
the deposition of the Betijoque Formation. The recycled gen published by Kohn et al. (1984) and Berm udez et al.

Table 3. Results of Carbonate content (CaCO3) and Rock-Eval analysis

Corg
CaCO3 Tmax S1 S2 TOC Ctot LECO Ctot LECO LECO HI (OSA) mg HI (LECO) mg
Sample (%) (°C) (mg g 1) (mg g 1) (%) (%) 1 (%) 2 (%) HC g 1 TOC HC g 1 TOC

Maracaibo
RH3 NA NA 0.0 0.0 0.1 0.1 0.1 NA NA NA
RH5 NA NA 0.0 0.0 0.0 0.0 0.0 NA NA NA
RV1 NA 400.0 3.2 16.5 51.0 50.9 53.3 NA 32.0 32.0
RV3 0.0 397.0 11.3 100.3 37.4 46.1 48.5 46.1 268.0 212.0
Barinas
CH52(1) NA 424 1.20 0.11 0.95 0.97 1.0 – 12 11
CH52(2) 38.5 389 0.88 0.41 4.74 8.9 9.3 4.51 9 9
PS39 NA NA 0.78 0.06 0.07 0.0 0.0

Note: NA indicates measure not available; Tmax is the maximum temperature; S1 indicates hydrocarbons concentration (gas and oil) present in the
rock; S2 indicates the amount of hydrocarbon produced from kerogen cracking. TOC is the weighted concentration of Carbon Organic Total expressed
in percentage; Ctot 1 and 2 correspond to the two measures of the weighted total carbon in percentage; Corg is the content of carbon organic; HI (OSA)
and HI (LECO) are measures of the Hydrogen Index.

© 2015 The Authors


18 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

(2010, 2011). Apatites with 10–15 Ma and 24–35 Ma However, no volcanic detritus has been observed in the
cooling ages are encountered in the Trujillo and Caparo Betijoque, Parangula and Rıo Yuca formations. There-
blocks, located towards the northeast and southwest fore, the youngest peak provides an upper limit to the
extremities of the Venezuelan Andes, respectively depositional age, given a certain lag time.
(Fig. 2; Bermudez et al., 2010). However, this relation only holds if fission tracks in
Given the location of our sections in the Maracaibo apatite were not reset after deposition because of burial
basin, the most likely source for these apatites would be heating, while a partial reset could alter the source-area
the Trujillo block (Fig. 2), although the Caparo source information. In our section there is no clear trend of
region could have been much closer to the section in Early increasing AFT (central or peak) ages with depth
Pliocene times than it is now, given the documented 30– (Fig 15a,b), as expected when AFT ages record source-
80 km right-lateral slip on the Bocono Fault since ~5 Ma area exhumation without thermal resetting (van der Beek
(Schubert, 1982; Audemard & Audemard, 2002). The et al., 2006). A slight decrease in central ages below
youngest peak ages at around 6  2 Ma are close to in- 2200 m depth of the section we sampled could indicate
situ ages encountered within the central Sierra La Culata, the onset of partial annealing in deeper samples
El Carmen and Sierra Nevada blocks and therefore record (Fig. 15b). However, the Hydrogen Index/Rock-Eval
the rapid exhumation of these central blocks of the orogen Tmax values of both samples plot below the temperature
in late Miocene – Early Pliocene times (Kohn et al., 1984; range needed for hydrocarbon formation (Espitalie et al.,
Berm udez et al., 2010, 2011). 1977), and the AFT total annealing zone. Therefore,
In addition to the biostratigraphic information pre- AFT ages were likely not significantly reset in our sam-
sented in this paper, detrital AFT ages provide an inde- ples and the youngest peak ages can be used as maximum
pendent estimate for the maximum depositional age, as values for depositional ages. These youngest-peak ages
the youngest grains or peak age in a detrital age spectrum cluster around 6  2 Ma for most samples.
should be equal to or older than the depositional age The fact that the sampled rocks were not sufficiently
(Kowallis et al., 1986; Bernet & Garver, 2005). In case of heated to significantly anneal fission tracks in apatite is
volcanic deposits, the AFT age is equal to the depositional not surprising. While paleo-geothermal gradients within
age, because of very rapid cooling following eruption. the Maracaibo basin are not well-constrained, it has been

Maracaibo basin Barinas basin


(a) 0 0
200 P1
200
P1
400 P2 400
P2
P3 600
600 P3
P4
Stratigraphic depth (m)
Stratigraphic depth (m)

P5 800
800
1000
1000
1200
1200
1400
1400
1600
1600
1800
1800 2000
2000 2200
2200 2400
2400 2600
2600 2800

0 10 20 30 40 50 0 10 20 30 40 50
Fission-track age (Ma) Fission-track age (Ma)

(b) Maracaibo basin Barinas basin


0 0
200 200
400 400
Central age 600 Central age
600
Stratigraphic depth (m)

Stratigraphic depth (m)

800
Palynological age (Ma)

Palynological age (Ma)

800
1000
1000
1200 6.3
1200 Presence of Stephanocolpites evansii
4.8 1400
1400 Cyatheacidites annulatus 10.4
(Lophosoria) 1600
1600 7.2 1800
1800 2000
2000 2200
2200 2400
2400 2600
2600 2800
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Fission-track age (Ma) Fission-track age (Ma)

Fig. 15. Upsection variation of AFT ages in the Maracaibo and Barinas basins. (a) Peak ages plotted against stratigraphic depth for
each age population (Pi). (b) Central ages plotted against stratigraphic depth. Independent constraints on stratigraphic ages are indi-
cated.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 19

M.A. Bermudez et al.

shown that foreland basins in general tend to be relatively broadly consistent with those presented by Erikson et al.
cool, with typical thermal gradients of 15–25 °C km 1 (2012) from two locations located along the same section.
(Ravenhurst et al., 1994; Allen & Allen, 2005; Coutand Their samples have similar central ages than ours (11.2–
et al., 2006), because of rapid sedimentation and infiltra- 12.7 Ma) but contain only a single grain-age component,
tion of meteoric water. Assuming a mean surface temper- because too few grains (6–20) were analysed to fully char-
ature of 20 °C and a thermal gradient of 20 °C km 1 even acterize the age spectrum. Our samples generally contain
the most deeply buried samples (2.7 km*20 °C km 1 significantly more grains, rendering recognition of differ-
+20 °C = 74 °C) in our sections would not have ent age components more effective, except for samples
exceeded 75 °C and therefore have hardly experienced PAN-1, PAN-3 and TPG-2. Apatites belonging to com-
the onset of partial annealing. Therefore, samples V97 ponents with 16–35 Ma peak ages are probably derived
and V95 of the Betijoque Formation indicate a maximum from the Trujillo or Caparo blocks. The 9–11 Ma age
Early Pliocene depositional age for the base of the section peak apatites were probably derived from the Sierra
we studied (Fig. 3, Table 2). Nevada and Sierra La Culata blocks, because in these
areas Middle-late Miocene AFT ages are encountered
today (Berm udez et al., 2010, 2011). The youngest peak
Barinas basin ages at around 5 Ma, detected only in one sample
(TPMY8) for Rıo Yuca Formation, should reflect
Sediment provenance
exhumation of the Santo Domingo gneiss on the southern
As in the Maracaibo basin, the Mio-Pliocene Barinas flank of the Venezuelan Andes (Fig. 2 in Berm udez et al.,
basin received erosional detritus from the crystalline core 2010); this 5 Ma peak age is also present in the modern
of the mountain belt, as shown by gneiss and granite lithic Santo Domingo River sediments (Fig. 2; Berm udez
grains in the sandstones (Fig. 9), The appearance of sedi- et al., 2013).
mentary lithic grains is an indication of sediment recy- Rock Eval analyses of sample CH52 from the Rıo Yuca
cling, particularly during deposition of the Parangula Formation (Fig. 14) show that burial heating of the sand-
Formation where no other lithic grains were observed. stones of this formation was insufficient for hydrocarbon
The recycled sedimentary rock fragments have three pos- formation. Therefore, AFT ages were not reset in samples
sible sources: (A) Mesozoic and Paleozoic sedimentary of the Rıo Yuca Formation. While the paleo-thermal gra-
cover rocks of the Venezuelan Andes, (B) Eocene shallow dient within the Barinas basin is not known, we can make
marine deposits, such as the fine-grained Pag€ uey Forma- the same estimate as for the Maracaibo basin. Using a typ-
tion or (C) in the upper Rio Yuca Formation sedimentary ical foreland-basin average thermal gradient of
clasts derived from the Middle to late Miocene sedimen- 20 °C km 1 and a surface temperature of 20 °C, even the
tary deposits. However, such grains were not found, are most deeply buried samples at about 2.7–3 km pale-
these deposits are in many places only poorly consolidated odepth of the section we studied have hardly experienced
and disintegrate easily during recycling. partial annealing, residing at temperature of <75–80 °C
for <10 Myr (Reiners & Brandon, 2006).
To better constrain the burial and exhumation history,
Implications of palynological results
we generated inverse models in order to study conver-
The analysis of pollen can also give information on sedi- gence of AFT data (ages, track lengths and kinetic indica-
ment recycling by the appearance of recycled paly- tors) and the available thermal maturity data (Erikson
nomorphs in younger sediments. Our pollen analyses of et al., 2012) of the geological formations, using the
upper Rıo Yuca Formation samples show abundant HeFTy program of Ketcham (2005) with the following
mature (dark brown) Eocene-Oligocene pollen and model parameters: we assumed that rocks were initially
dinoflagellates, a clear indication for the recycling of exhumed between 15 and 11 Ma, buried in the foreland
Eocene sediments of the Pag€ uey Formation. Despite the basin between ~10 and 5 Ma, held at peak temperatures
recycling of older pollen, the presence of biostratigraphic between 60 and 140 °C for an interval of ~5 My, and
marker species form a clear indication of the late Mio- exhumed to the surface between 3–1 Ma and the present.
cene-Early Pliocene depositional age (<7.15–3.40 Ma) of The modelling results of the best-constrained samples
the formation. In contrast with the Maracaibo basin, no (e.g. TPG-1, TPMY-5, TPMY-4) suggest a maximum
pollen species indicative of high-altitude vegetation were burial temperature between not exceeding 90 °C, and
encountered; the samples rather record the low-lying probably around 75 °C, between ~6 and ~1 Ma (Fig. 16);
tropical forest environment of the depositional site. predicted vitrinite reflectance values between 0.57–0.62
(using the Easy-Ro model; Sweeney & Burnham, 1990).
This finding is consistent with vitrinite reflectance values
Detrital thermochronology
of 0.56–0.67 for samples from the Rıo Yuca Formation
Similar to the Maracaibo basin, the detrital AFT ages in reported by Erikson et al. (2012), as well as with their
the Barinas basin correspond closely to the published bed- thermal modelling of an AFT sample from the base of this
rock AFT ages of the Venezuelan Andes (Kohn et al., formation, which shows peak burial temperatures of
1984; Berm udez et al., 2010, 2011). The data are also ~80 °C. Therefore, a maximum Middle – late Miocene

© 2015 The Authors


20 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

depositional age (<12 Ma) for the Parangula Formation est age peak of the Maracaibo basin samples. While this
can be discerned from sample TPG-2, in the lower part of difference is probably partly due to the older sedimentary
the studied section (Fig. 5, Table 2) and which has a sin- record in the Barinas basin, it also appears in samples that
gle age population of 14  3 Ma (assuming short lag should have similar Early Pliocene apparent cooling ages.
times between exhumation and deposition on the order of Unfortunately, the lack of precise depositional ages of the
~2 My). Erikson et al. (2012) reached a similar conclusion sampled sandstone units in both basins prohibits a more
for the maximum age of these deposits. The AFT ages of detailed comparison.
samples PAN-3 and PAN-1 provide evidence for
exhumation-driven cooling of the Pag€ uey Formation and
Neogene surface uplift and exhumation of
the lower part of the Parangula Formation starting at
theVenezuelan Andes in comparison to the
about 6–5 Ma. Therefore, apatite fission tracks in the
Eastern Cordillera of Colombia
Pag€uey Formation were totally reset and the observed
AFT cooling ages provide information about the timing The development of the Northern Andes has attracted
of exhumation of this formation. Unfortunately no track- significant attention in recent years. It has long been rec-
lengths data are available for time-temperature history ognized that the formation of the Western and Central
modelling of these samples. Cordillera in Colombia and Ecuador are related to the
The static peaks observed in the detrital samples from accretion of oceanic terranes during the Late Cretaceous
both foreland-basin sections (Fig. 15) imply episodic to Paleocene (e.g. Case et al., 1990; Dengo & Covey,
exhumation of the central tectonic blocks in the Venezue- 1993; Gregory-Wodzicki, 2000; Gomez et al., 2005; Jail-
lan Andes (cf. Berm udez et al., 2010, 2011). A static peak lard et al., 2009). Within the Eastern Cordillera of
is an age peak that remains fairly constant up-section in Colombia, Cretaceous quartzose sandstones and
the stratigraphic record, which means the lag time mudrocks are widely exposed, and to a lesser extent Juras-
increases (e.g. Garver & Brandon, 1994a,b). In contrast, sic graben conglomerates and Paleocene to Eocene non-
continuous exhumation would provide a ‘moving’ peak, marine sandstones. Paleogene sediments are derived from
which becomes younger upsection (e.g. Garver & Bran- the erosion of the Central Cordillera to the west and were
don, 1994a,b). While one could argue that deposition of deposited prior to uplift of the Eastern Cordillera along
the Betijoque Formation was very rapid (~3–1.5 Ma), the reactivated former graben structures (e.g. Dengo &
Parangula and Rıo Yuca formations represent up to Covey, 1993; Corredor, 2003; Mora et al., 2009, 2010b;
10 Ma of deposition, a sufficient time span to discrimi- Horton et al., 2010b).
nate a moving from a static AFT age peak (e.g. van der The surface-uplift history of the Eastern Cordillera is
Beek et al., 2006). still a matter of debate (Horton et al., 2010b). While Bay-
Interestingly, the youngest age peak in the Barinas ona et al. (2008, 2013) provide sedimentological evidence
basin samples tends to be 3–5 Myr older than the young- for five episodes of deformation during the evolution of

Time-temperature history: TPMY-5 (1700 m, n = 48)


0
AFT track-length distribution

20 0.60

0.50
40
Temperature (°C)

0.40
Frequency

60

0.30
80
0.20
100
Obs. age: 12.3 Ma; Mod. Age: 12.1 Ma; 0.10
Obs. MTL: 13.4 μm; Mod. MTL: 13.8 μm;
120 GOF Age: 0.66; GOF MTL: 0.73;
Obs. % Ro: 0.67; Mod. % Ro: >0.41; 0.00
GOF % Ro: 0.99 0 2 4 6 8 10 12 14 16 18 20
140 Length (µm)
15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 0
Time (Ma)

Fig. 16. Time-temperature (t-T) history of sample TPMY-5 as reconstructed using the AFT data (central age, track-length distribu-
tion and kinetic indicator Dpar) and HeFTy inversion code (Ketcham, 2005). Thick black and blue lines represent overall best-fit ther-
mal history, and weighted mean t-T path; purple and green ranges show thermal histories that provide a ‘good’ and ‘acceptable’ fit to
the data, respectively, as defined by Ketcham (2005). Fit to track-length distribution is shown in the track-length histogram to the
right.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 21

M.A. Bermudez et al.

the Eastern Cordillera, other authors prefer a more con- Andes are more or less contemporaneous but not driven
tinuous deformation history, with migration of the defor- by exactly the same underlying geodynamic processes.
mation front toward the east. Most authors agree that While both ranges are doubly vergent, thick-skinned
early deformation in the area of the Eastern Cordillera orogenic wedges that have similar modern mean eleva-
may have started in the Late Cretaceous (subsidence and tions, the Eastern Cordillera is roughly twice as wide as
foreland basin formation), but was definitely underway in the Venezuelan Andes (~100 km). The Eastern Cordil-
mid-late Eocene to Late Oligocene times (e.g. Corredor, lera is more or less N-S oriented, while the Venezuelan
2003; Horton et al., 2010a,b; Mora et al., 2010a,b; Parra Andes are largely NE-SW oriented. The Venezuelan
et al., 2010; Ochoa et al., 2012), with onset of surface Andes contain their highest relief in the central zone
uplift in the axial zone (26–23 Ma; Horton et al., 2010b). along the Bocono strike-slip fault. The axial zone of the
Structural and geochronologic data from detrital zircon Eastern Cordillera, in contrast, is the high plateau of the
from both sides and across the Eastern Cordillera imply Sabana de Bogota with much less relief (Mora et al.,
that the major phase of surface uplift and associated ero- 2008). An interesting difference between the two moun-
sional exhumation started in the Late Oligocene to early tain belts is that exposure of crystalline basement rocks
Miocene (e.g.; Horton et al., 2010a,b; Nie et al., 2010). is much more widespread in the Venezuelan Andes than
Mora et al. (2010c) suggested that plate deformation had in the Eastern Cordillera. In addition, the Venezuelan
started well before the late Miocene in the Colombian Andes appear more symmetric in deformation style and
Eastern Cordillera, but no evidence for high paleoeleva- bedrock exposure than the asymmetric Eastern Cordil-
tion exists before the late Miocene. Based on pollen data lera (Mora et al., 2008). The link between both orogens,
by Van der Hammen et al. (1973) and Hooghiemstra the Santander Massif and the Santa Marta-Bucaramanga
et al. (2006), Mora et al. (2010c) favour surface uplift of fault (SBF; Fig. 1) represent key elements to understand
the Colombian Eastern Cordillera towards present-day the evolution of this area, because they correspond to
elevations between 15 and 3 Ma. Bedrock AFT data from the NE termination of the ‘classical’ Colombian/
the eastern flank range from 1 to 3 Ma across the Que- Ecuadorian Andes with the Western, Central and East-
tame massif (Mora et al., 2008; Parra et al., 2009a,b) and ern Cordilleras and the seemingly independent Venezue-
other parts of the Eastern Cordillera (Mora et al., 2010b), lan Andes.
while the deposition of thick Miocene-Pliocene sediments It still needs to be determined how the surface uplift
in the foreland basin to the east of the Eastern Cordillera of the Eastern Cordillera in Colombia and the Venezue-
(e.g. Gomez et al., 2005; Bayona et al., 2008, 2013; Parra lan Andes influenced the local and regional climate. It is
et al., 2009b, 2010; Horton et al., 2010b) demonstrates obvious from the bedrock AFT data that Pliocene ero-
late Miocene to Pleistocene accelerated erosional exhuma- sional exhumation is focused on the currently relatively
tion and possible surface uplift of the Eastern Cordillera. wet eastern flank of the Eastern Cordillera, where
Gregory-Wodzicki (2000) proposed that from the Middle 1–3 Ma cooling ages are observed (Mora et al., 2008,
Miocene through the Early Pliocene, the Eastern Cordil- 2010b; Parra et al., 2009a,b), while such young AFT
lera only reached about 40% of its present-day mean ele- cooling ages in the Venezuelan Andes have so far only
vation and that surface uplift between 5 and 2 Ma been observed in outcrops of the relatively dry central
brought the mountain belt to its current elevation. This zone along the Bocono strike slip fault (Kohn et al.,
timing coincides with the final closure of the Panama isth- 1984; Berm udez et al., 2010). Therefore, it has been pro-
mus and the end of the collision of the Panama arc with posed that Pliocene exhumation of the Venezuelan Andes
northwestern South America (Bartoli et al., 2005; Farris is principally controlled by tectonics (Berm udez et al.,
et al., 2011). 2013), while climate could have a larger role in driving
As shown by the detrital AFT data of the present exhumation of the Colombian Eastern Cordillera (Mora
study and the published bedrock and detrital AFT data et al., 2008).
from the Venezuelan Andes (Kohn et al., 1984; Shagam
et al., 1984; Berm udez et al., 2010, 2011, 2013), the
main phase of exhumation in the Venezuelan Andes
CONCLUSIONS
started in the Middle-late Miocene and continued into
the Pliocene. The palynological data presented here pro- Sediment petrographic, palynologic and thermochrono-
vide evidence that the mountain belt had reached its logic evidence from the pro- and retro-wedge foreland
present-day elevation during the late Miocene to Plio- basins of the Venezuelan Andes lead us to the following
cene, apparently somewhat earlier than the Colombian conclusions:
Eastern Cordillera. The clockwise rotation of the Mara-  The sediment petrography indicates erosion of crys-
caibo block and the collision of the Panama arc with talline (granitic and gneissic) bedrock in the core of
northwestern South America are regarded as the driving the mountain belt, as well as recycling of sedimentary
forces responsible for the surface uplift of the individual cover units on its flanks, at least since the late Mio-
tectonic blocks that make up the Venezuelan Andes cene. The basins on both sides of the orogen received
(Berm udez et al., 2010, 2011, 2013). Therefore, the evo- sediments of roughly similar composition, which are
lution of the Eastern Cordillera and the Venezuelan characterized by a coarsening upward trend.

© 2015 The Authors


22 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

 Contrary with previous estimates, the fluvial Paran- Andean palynology) for their observations on taxonomy
gula Formation in the Barinas basin appears to be of and interpretation, Andres Pardo for facilitating the
Middle-late Miocene age, based on our AFT and pol- microphotography and Jan van Arkel for mounting the
len data, instead of Oligocene-Middle Miocene. plates. Francis Coeur and Francßois Senebier are thanked
 Rock-Eval analyses of organic matter from the Betijo- for sample preparation and mineral separation. We thank
que and Rıo Yuca formations indicate that these units Isabelle Coutand, Barry Kohn, Massimiliano Zattin and
were not sufficiently heated for kerogen production. Basin Research editor Andreas Mulch for constructive
Therefore, the burial temperatures were insufficient criticism that allowed to significantly improve the manu-
for resetting of fission tracks in apatite. As a conse- script.
quence, the detrital AFT age patterns from the appar-
ently unreset samples mainly reflect source rock
cooling.
 Reset AFT ages of 6–5 Ma of samples from the
CONFLICT OF INTEREST
Eocene Pag€ uey Formation and the very base of the No conflict of interest declared.
Parangula Formation indicate exhumational cooling
on the southern flank of the Venezuelan Andes in the
latest Miocene. This is consistent with the observation
SUPPORTING INFORMATION
of sediment recycling of Eocene sedimentary rocks
and an increasing number of sedimentary rock frag- Additional Supporting Information may be found in the
ments upsection in the Barinas basin. online version of this article:
 The detrital AFT data presented in this work show
relatively static minimum ages in each sample at about Tables S1. Samples and pollen morphological taxa
5  2 Ma in the Pliocene Maracaibo basin sediments found at the Rı́o Yuca Formation, Barinas-Apure Basin.
and about 7  2 Ma in the Mio-Pliocene Barinas Tables S2. Samples and pollen morphological taxa
basin sediments. These data are consistent with the found at the Betijoque Formation, Maracaibo Basin.
bedrock AFT ages in the Venezuelan Andes of Kohn
et al. (1984) and Berm udez et al. (2010, 2011) and
suggest rapid episodic exhumation of different fault-
bounded blocks within this mountain belt.
 The palynological data show that the Venezuelan REFERENCES
Andes had reached their present-day elevation (3500– ALLEN, P.A. & ALLEN, J.R. (2005) Basin Analysis: Principles and
4000 m) by the late Miocene – Pliocene. In addition, Applications, 2nd ed. Blackwell Publishing, Oxford, UK. 549
the recycling of older pollen supports the notion of pp.
sediment recycling based on petrography. AUDEMARD, F.E. & AUDEMARD, F.A. (2002) Structure of the
 The main phase of surface uplift, creation of impor- Merida Andes, Venezuela: relations with the South America-
tant topography and relief, and erosional exhumation Caribbean geodynamic interaction. Tectonophysics, 345, 299–
in the Venezuelan Andes occurred during the late 327.
Miocene to Pliocene. Therefore, the evolution of the BARTOLI, G., SARNTHEIN, M., WEINELT, M., ERLENKEUSER, H.,

GARBE-SCHONBERG , G. & LEA, D.W. (2005) Final closure of
Eastern Cordillera in Colombia and the Venezuelan
Panama and onset of nothern hemisphere galciation. Earth
Andes were not fully contemporeanous. The two
Planet. Sci. Lett., 237, 33–44.
mountain belts are considerably different in size, BAUDIN, F., TRIBOVILLARD, N. & TRICHET, J. (2007) Geologie de
structure and exhumation history. la Matiere Organique, Edition Vuibert, Paris, Collection
Interactions, 263 pp.
BAYONA, G., CORTES  , M., JARAMILLO, C., OJEDA, G., ARISTIZA-
BAL, J.J. & HARKER, A.R. (2008) An integrated analysis of an

ACKNOWLEDGEMENTS orogen-sedimentary basin pair: latest Cretaceous-Cenozoic


evolution of the linked Eastern Cordillera orogen and the Lla-
This study was supported by the Consejo de Desarrollo nos foreland basin of Colombia. Geol. Soc. Am. Bull., 120,
Cientıfico y Humanıstico (CDCH-UCV) de la Universi- 1171–1197.
dad Central de Venezuela (UCV), Projects PI09-00-6219- BAYONA, G., CARDONA, A., JARAMILLO, C., MORA, A., MONTES,
2006, PG 08-8273-2011, Fondo Nacional de Ciencia y C., CABALLERO, V., MAHECHA, H., LAMUS, F., MONTENEGRO,
Tecnologıa (FONACIT) Project Number S1-200000636, O., JIMENEZ, G., MESA, A. & VALENCIA, V. (2013) Onset of
GIAME Project provided to FUNVISIS, ECOS Nord fault reactivation in the Eastern Cordillera of Colombia and
proximal Llanos Basin; response to Caribbean–South Ameri-
project V08U01 and support provided by Petroleos de
can convergence in early Palaeogene time. In: Thick-Skin-
Venezuela, S.A (PDVSA) in 1990–1995. We thank Dominated Orogens: From Initial Inversion to Full Accretion
Millerlandy Romero-Baez (pollen/biostratigraphy), Roel (Ed. by M. Nemcok, A.R. Mora & J.W. Cosgrove), pp. 377.
Verreussel (dinoflagellates/biostratigraphy), Zaire Geological Society, Special Publications, London, 10.1144/
Gonzalez-Carranza and Giovanni Bogota (Quaternary SP377.5.

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 23

M.A. Bermudez et al.

van der BEEK, P.A., ROBERT, X., MUGNIER, J.L., BERNET, M., ERIKSON, J.P., KELLEY, S.A., OSMOLOVSKY, P. & VEROSUB, K.L.
HUYGHE, P. & LABRIN, E. (2006) Late Miocene – Recent (2012) Linked basin sedimentation and orogenic uplift: the
exhumation of the central Himalaya and recycling in the fore- Neogene Barinas basin sediments derived from the Venezue-
land basin assessed by apatite fission-track thermochronology lan Andes. J. S. Am. Earth Sci., 39, 138–156.
of Siwalik sediments, Nepal. Basin Res., 18, 413–434. ESCALONA, A. & MANN, P. (2006) An overview of the petroleum

BERMUDEZ , M.A., KOHN, B., van der BEEK, P., BERNET, M., system of Maracaibo basin. Am. Assoc. Pet. Geol. Bull., 90,
O’SULLIVAN, P. & SHAGAM, R. (2010) Spatial and temporal 657–678.
patterns of exhumation across the Venezuelan Andes: impli- ESPITALIE , J., LAPORTE, J.L., MADEC, M., MARQUIS, F., LEPLAT,
cations for Cenozoic Caribbean geodynamics. Tectonics, 29, P., PAULET, J. & BOUTEFEU, A. (1977) Rapid method for
TC5009. source rock characterization, and for determination of their

BERMUDEZ , M.A., van der BEEK, P. & BERNET, M. (2011) Asyn- petroleum potential and degree of evolution. Revue de l’Insti-
chronous Mio-Pliocene exhumation of the central Venezuelan tut Francais du Petrole et Annales des Combustibles Liquides, 32
Andes. Geology, 39, 139–142. (1), 23–42.

BERMUDEZ , M.A., van der BEEK, P. & BERNET, M. (2013) Strong ESPITALIE , J., DEROO, G. & MARQUIS, F. (1986) La pyrolyse Rock
tectonic and weak climatic control on exhumation rates in the Eval et ses applications. Rev. Inst. Fr. Petr, 32(1) 23–42.
Venezuelan Andes. Lithosphere, 5, 3–16. FARRIS, D.W., JARAMILLO, C.A., BAYONA, G.A., RESTREPO-MOR-

BERMUDEZ , M.A., GLOTZBACH, C. & ALSON, P. (2015) A new ENO, S.A., MONTES, C., CARDONA, A., MORA, A., SPEAKMAN,
Poissonian algorithm for the determination of fission-track R.J., GLASSCOCK, M.D., REINERS, P. & VALENCIA, V. (2011)
ages. Computers & Geosciences, 76, 141–150. Fracturing of the Panamanian Isthmus during initial collision
BERNET, M. & GARVER, J.I. (2005) Fission-track analysis of with South America. Geology, 39(11), 1007–1010.
detrital zircon. In: Low-temperature Thermochronology, FOLK, R.L. (1965) Petrology of Sedimentary Rocks. Hemphill
Reviews in Mineralogy and Geochemistry (Ed. by P. Reiners & Publishing Company, Austin, Texas, 190 pp.
T. Ehlers), 58, 205–237. GALBRAITH, R.F. & LASLETT, G.M. (1993) Statistical models for
BRINKHUIS, H. & BIFFI, U. (1993) Dinoflagellate cysts stratigra- mixed fission track ages. Nucl. Tracks Radiat. Meas., 21, 459–
phy of the Eocene/Oligocene transition in central Italy. Mar. 470.
Micropalaeontol., 22, 131–183. GARVER, J.I. & BRANDON, M.T. (1994a) Fission track ages of
BUJAK, J.P. & MUDGE, D. (1994) A high-resolution North Sea detrital zircons from Cretaceous strata, southern British
dinocyst zonation. J. Geol. Soc. Lond., 151, 449–462. Columbia: implications for the Baja BC hypothesis. Tectonics,
CASE, J.E., SHAGAM, R. & GIEGENGACK, R.F. (1990) Geology of 13(2), 401–420.
the Northern Andes; an overview. In: The Caribbean Region: GARVER, J.I. & BRANDON, M.T. (1994b) Erosional denudation of
The Geology of North America, H: Boulder, CO. (Ed. by the British Columbia Coast Ranges as determined from fission
Dengo G. & Case J.E.) Geol. Soc. Am., 177–200. track ages of detrital zircon from the Tofino Basin, Olympic

CEDIEL, F., SHAW, R.P. & CACERES , C. (2003) Tectonic assembly Peninsula, Washington. Geol. Soc. Am. Bul., 106(11), 1398–
of the Northern Andean Block, in: The Circum-Gulf of Mexico 1412.
and the Caribbean: hydrocarbon Habitats, Basin Formation, and GARZANTI, E., VEZZOLI, G., ANDO , S., LAVE , J., ATTAL, M.,
Plate Tectonics. (Ed. by Bartolini C, Buffler R.T. & Blickwede FRANCE-LANORD, C. & DECELLES, P. (2007) Quantifying sand
J.) AAAPG Mem., 79, 815–848. provenance and erosion (Marsyandi River, Nepal Himalaya).

CHACIN, L., JACOME , M.I. & IZARRA, C. (2005) Flexural and Earth Planet. Sci. Lett., 258, 500–515.
gravity modelling of the Merida Andes and Barinas-Apure GERMERAAD, J.H., HOPPING, C.A. & MULLER, J. (1968) Palynol-
Basin, Western Venezuela. Tectonophysics, 405, 155–167. ogy of Tertiary sediments from tropical areas. Rev. Palaeobot.
COLLETTA, B., ROURE, F., DE TONI, B., LOUREIRO, D., PAS- Palynol., 6, 189–348.
SALACQUA, H. & GOU, Y. (1997) Tectonic inheritance, crustal 
GOMEZ , E., JORDAN, T.E., ALLMENDINGER, R.W. & CARDOZO, N.
architecture, and contrasting structural styles in the Venezue- (2005) Development of the Colombian foreland-basin system
lan Andes. Tectonics, 16, 777–794. as a consequence of diachronous exhumation of the northern
COLMENARES, L. & ZOBACK, M.D. (2003) Stress field and seis- Andes. Geol. Soc. Am. Bull., 117, 1272–1292.
motectonics of northern South America. Geology, 31, 721– 
GONZALEZ -GUZMAN  , A.E. (1967) A palynological study on the
724. Upper Los Cuervos and Mirador Formations (Lower and Middle
CORREDOR, F. (2003) Seismic strain rates and distributed conti- Eocene; Tibu area, Colombia). E.J. Brill, Leiden, 68 pp.
nental deformation in the northern Andes and three-dimen- GREGORY-WODZICKI, K.M. (2000) Uplift history of the Central
sional seismotectonics of the northwestern South America. and Northern Andes: a review. Geol. Soc. Am. Bull., 112,
685. Tectonophysics, 372, 147–166. 1091–1105.
COUTAND, I., CARRAPA, B., DEEKEN, A., SCHMIDT, A.K., SOBEL, GUERRERO, O., JIMENEZ  , D., LAYA, J., MONSALVE, M. &
E.R. & STRECKER, M.R. (2006) Propagation of orographic bar- ALVARADO, M. (2009) Sedimentation of Miocene Upper-
riers along an active range front: insights from sandstone pet- Pliocene in the north flank of the central Venezuelan
rography and detrital apatite fission-track thermochronology Andes. Revista Ciencia e Ingenierıa, 30(1), 47–56. ISSN
in the intramontane Angastaco basin, NW Argentina. Basin 1316-7081.
Res., 18, 1–26. GUZMAN  , J.I. & FISHER, W.L. (2006) Early and middle Miocene
DE TONI, B. & KELLOGG, J. (1993) Seismic evidence for blind depositional history of the Maracaibo basin, western Vene-
thrusting of the Northwestern flank of the Venezuelan Andes. zuela. Am. Assoc. Pet. Geol. Bull., 90, 625–655.
Tectonics, 12, 1393–1409. HACKLEY, P.C., URBANI, F., KARLSEN, A.W. & GARRITY, C.P.
DENGO, C.A. & COVEY, M.C. (1993) Structure of the Eastern (2005) Geologic shaded relief map of Venezuela. U.S. Geo-
Cordillera of Colombia: implications for trap styles and regio- logical Survey Open File Report 2005–1038. (Available at:
nal tectonics. Am. Assoc. Pet. Geol. Bull., 77, 1315–1337. http://pubs.usgs.gov/of/2005/1038/).

© 2015 The Authors


24 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists
Surface uplift and exhumation in basins of the VA

HENRIQUES-CASAS, L.J. (2004) Modelo geodinamico integrado KOWALLIS, B.J., HEATON, J.S. & BRINGHURST, K. (1986) Fission-
de la Cuenca barinas–apure y los Andes de merida, estado track dating of volcanically derived sedimentary rocks. Geol-
barinas. Master Thesis, Universidad Simon Bolıvar. ogy, 14, 19–22.
HOOGHIEMSTRA, H. (1984) Vegetational and climatic history of LORENTE, M.A. (1986) Palynology and palynofacies of the
the high plain of Bogota, Colombia: a continuous record of Upper Tertiary in Venezuela. Dissertationes Botanicae, 99,
the last 3.5 million years. Dissertationes Botanicae, 79, 1–368. 222 pp.
J. Cramer Verlag, Vaduz. (Also published as van der Ham- LUGO, J. & MANN, P. (1995) Jurassic-Eocene tectonic evolution
men, T., ed., The Quaternary of Colombia, v. 10 Amster- of maracaibo Basin, Venezuela. In: Petroleum Basins of South
dam). America (Ed. by Tankard A.J., Suarez R.S. & Welsink H.J.)
HOOGHIEMSTRA, H., WIJNINGA, V.M. & CLEEF, A.M. (2006) Am. Assoc. Pet. Geol. Mem., 62, 699–725.
The Paleobotanical record of Colombia: implications for MONTES, C., HATCHER, R.D. & RESTREPO-PACE, P. (2005) Tec-
biogeography and biodiversity. Ann. Mo. Bot. Gard., 93, tonic reconstruction of the Northern Andean blocks: oblique
297–325. convergence and rotations derived from the kinematics of the
HOORN, C., GUERRERO, J., SARMIENTO, G.A. & LORENTE, M.A. Piedras-Girardot area, Colombia. Tectonophysics, 399, 221–
(1995) Andean tectonics as a cause for changing drainage pat- 250.
terns in Miocene northern South America. Geology, 23, 237– MONTES, C., GUZMAN  , G., BAYONA, G., CARDONA, A., VALENCIA,
240. V. & JARAMILLO, C. (2010) Clockwise rotation of the Santa
HOORN, C., WESSELINGH, F.P., TER STEEGE, H., BERMUDEZ  , Marta Massif and simultaneous Paleogene to Neogene defor-

M.A., MORA, A., SEVINK, J., SANMARTIN, I., SANCHEZ-ME- mation of the Plato-San Jorge and Cesar-Rancherıa Basins. J.
SEGUER, A., ANDERSON, C.L., FIGUEIREDO, J.P., JARAMILLO, S. Am. Earth Sci., 29, 832–848.
C., RIFF, D., NEGRI, F.R., HOOGHIEMSTRA, H., LUNDBERG, J., MORA, A., PARRA, M., STRECKER, M.R., SOBEL, E.R.,

STADLER, T., SARKINEN , T. & ANTONELLI, A. (2010) Amazo- HOOGHIEMSTRA, H., TORRES, V. & JARAMILLO, J.V. (2008) Cli-
nia through time: andean uplift, climate change, landscape matic forcing of asymmetric orogenic evolution in the Eastern
evolution, and biodiversity. Science, 330, 927–931. Cordillera of Colombia. Geol. Soc. Am. Bull., 120, 930–949.
HORTON, B.K., PARRA, M., SAYLOR, J.E., NIE, J., MORA, A., MORA, A., GOANA, T., KLEY, J., MONTOYA, D., PARRA, M.,
TORRES, V., STOCKLI, D.F. & STRECKER, M.R. (2010a) Re- QUIROZ, L.J., REYES, G. & STRECKER, M.R. (2009) The role of
solving uplift of the northern Andes using detrital zircon age inherited extensional fault segmentation and linkage in con-
signatures. GSA Today, 20, 4–9. tractional orogenesis: a reconstruction of Lower Cretaceous
HORTON, B.K., SAYLOR, J.E., NIE, J., MORA, A., PARRA, M., inverted rift basins in the Eastern Cordillera of Colombia.
REYES-HARKER, A. & STOCKLI, D.F. (2010b) Linking sedi- Basin Res., 21, 111–137.
mentation in the northern Andes to basement configuration, MORA, A., PARRA, M., STRECKER, M.F., SOBEL, E.R., ZEILINGER,
Mesozoic extension, and Cenozoic shortening: evidence from G., JARAMILLO, C., DA SILVA, S.F. & BLANCO, M. (2010a)
detrital zircon U-Pb ages, Eastern Cordillera, Colombia. Geol. The eastern foothills of the Eastern Cordillera of Colombia:
Soc. Am. Bull., 122, 1423–1442. an example of multiple factors controlling structural styles

JACOME , M. & CHACIN, L. (2008) Subsidence history of the Bari- and active tectonics. Geol. Soc. Am. Bull., 122, 1846–1864.
nas Apure Basin: Western Venezuela. Back to Exploration: MORA, A., HORTON, B.K., MESA, A., RUBIANO, J., KETCHAM,
2008 CSPG CSEG CWLS Convention Calgary, Canada, R.A., PARRA, M., BLANCO, V., GARCIA, D. & STOCKLI, D.F.
682–686. (2010b) Migration of Cenozoic deformation in the Eastern
JAILLARD, E., LAPIERRE, H., ORDONEZ, M., TORO ALAVA, J., Cordillera of Colombia interpreted from fission-track results
AMORTEGUI, A. & VAN MELLE, J. (2009) Accreted oceanic ter- and structural relationships: implications for petroleum sys-
ranes in Ecuador: southern edge of the Caribbean Plate? In tems. Am. Assoc. Pet. Geol. Bull., 94, 1543–1580.
(Ed. by K.H. James, M.A. Lorente & J.L. Pindell). The Origin MORA, A., BABY, P., RODDAZ, M., PARRA, M., BRUSSET, S., HER-
and Evolution of the Caribbean Plate. Geological Society of MOZA, W. & ESPURT, N. (2010c), Tectonic history of the
London, Special Publication, London, 328, 469–485. Andes and sub-Andean zones: implications for the develop-
JAMES, H.K. (2000) The Venezuelan hydrocarbon habitat, part ment of the Amazon drainage basin. In: Amazonia: Landscape
1: tectonics, structure, palaeogeography and source rocks. J. and Species Evolution: A Look into the Past (Ed. by C. Hoorn
Pet. Geol, 23(1), 5–53. & F.P. Wesselingh), pp. 38–60. Wiley-Blackwell, Chichester,
JARAMILLO, C. & RUEDA, M. (2008) A morphological UK.
electronic database of Cretaceous-Tertiary fossil pollen and MULLER, J., DI GIACOMO, E. & VAN ERVE, A.W. (1987) A paly-
spores from northern South America. Bucaramanga: nological zonation for the Cretaceous, Tertiary, and Quater-
Smithsonian Tropical Research Institute and Colombian nary of Northern South America. AASP Contrib. Series, 19,
Petroleum Institute (http://biogeodb.stri.si.edu/jaramillo/ 7–76.
palynomorph/). NIE, J., HORTON, B.K., MORA, A., SAYLOR, J.E., HOUSH, T.B.,
JARAMILLO, C.A., RUEDA, M. & TORRES, V. (2011) A palynologi- RUBIANO, J. & NARANJO, J. (2010) Tracking exhumation of
cal zonation for the Cenozoic of the Llanos and Llanos Foot- Andean ranges bounding the Middle Magdalena Valley Basin,
hills of Colombia. Palynology, 35(1), 46–84. Colombia. Geology, 38, 451–454.
KETCHAM, R.A. (2005) Forward and inverse modelling of low- OCHOA, D., HOORN, C., JARAMILLO, C., BAYONA, G., PARRA, M.
temperature thermochronology data. Rev. Mineral. Geochem., & DE LA PARRA, F. (2012) The final phase of tropical lowland
58, 275–314. conditions in the axial zone of the Eastern Cordillera of
KOHN, B.P., SHAGAM, R., BANKS, P.O. & BURKLEY, L.A. (1984) Colombia: evidence from three palynological records. J. S.
Mesozoic-Pleistocene fission track ages on rocks of the Am. Earth Sci., 39, 157–169.
Venezuelan Andes and their tectonic implications. Geol. Soc. PARRA, M., MORA, A., JARAMILLO, C., STRECKER, M.R., SOBEL,
Am. Mem., 162, 365–384. E.R., QUIROZ, L., RUEDA, M. & TORRES, V. (2009a) Orogenic

© 2015 The Authors


Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists 25

M.A. Bermudez et al.

wedge advance in the northern Andes: evidence from the SHAGAM, R., KOHN, B.P., VARGAS, R., RODRIGUEZ, G.I.,
Oligocene-Miocene sedimentary record of the Medina Basin, BANKS, P.O., DASCH, L.E. & PIMENTEL, N. (1984)
Eastern Cordillera, Colombia. Geol. Soc. Am. Bull., 121, 780– Tectonic implications of Cretaceous-Pliocene fission-track
800. ages from rocks of the circum-Maracaibo Basin region of
PARRA, M., MORA, A., SOBEL, E.R., STRECKER, M.R. & GON- western Venezuela and eastern Colombia. Geol. Soc. Am.

ZALEZ , R. (2009b) Episodic orogenic front migration in the Mem., 162, 385–412.
northern Andes: constrains from low-temperature ther- STEWART, R.J. & BRANDON, M.T. (2004) Detrital zircon fission-
mochronology in the Eastern Cordillera, Colombia. Tectonics, track ages for the “Hoh Formation”: implications for late
28,TC4004, 1–27. doi:10.1029/2008TC002423. Cenozoic evolution of the Cascadia subduction wedge. Geol.
PARRA, M., MORA, A., JARAMILLO, C., TORRES, V., ZEILINGER, G. Soc. Am. Bull., 116, 60–75.
& STRECKER, M. (2010) Tectonic controls on Cenozoic fore- SWEENEY, J.J. & BURNHAM, A.K. (1990) Evaluation of a simple
land basin development in the north-eastern Andes, Colom- model of vitrinite reflectance based on chemical kinetics. Am.
bia. Basin Res., 22, 874–903. Assoc. Pet. Geol. Bull., 74(10), 1559–1570.
PINDELL, J.L. & KENNAN, L. (2009) Tectonic evolution of the TORRES-TORRES, V. (2006) Pliocene-Pleistocene evolution of
Gulf of Mexico, Caribbean and northern South America in flora, vegetation and climate: A palynological and sedimento-
the mantle reference frame: an update. In: The Origin and logical study of 586-m core from the Bogota Basin, Colombia.
Evolution of the Caribbean Plate (Ed. by James K.H., Lorente Ph. D. Thesis, University of Amsterdam, The Netherlands,
M.A. & Pindell J.L.). Geological Society London Special 181 p and color plates.
Publications, 328, 1–55. VAN DER HAMMEN, T., WERNER, J.H. & DOMMELEN, V. (1973)
POWELL, A.J. (1992) Dinoflagellate cysts of the Tertiary System. Palynological record of the upheaval of the Northern Andes: a
In: A Stratigraphic Index of Dinoflagellate Cysts (Ed. by A.J. study of the Pliocene and Lower Quaternary of the Colom-
Powell), pp. 155–251. Chapman and Hall, London. bian Eastern Cordillera and the early evolution of its High-
RAVENHURST, C.E., WILLETT, S.D., DONELICK, R.A. & BEAU- Andean biota. Rev. Palaeobot. Palynol., 16, 1–122.
MONT, C. (1994) Apatite fission track thermochronometry VAN’T VEER, R. & HOOGHIEMSTRA, H. (2000) Montane forest
from central Alberta: implications for the thermal history of evolution during the last 650 000 yr in Colombia: a multivari-
the Western Canada Sedimentary Basin. J. Geophys. Res., 99, ate approach based on pollen record Funza-I. J. Quat. Sci.,
20023–20042. 15, 329–346.
REINERS, P.W. & BRANDON, M.T. (2006) Using thermochronol- VERMEESCH, P. (2004) How many grains are needed for a prove-
ogy to understand orogenic erosion. Ann. Rev. Earth Planet. nance study? Earth Planet. Sci. Lett., 224, 441–451.
Sci., 34, 419–466. 
ZAMBRANO, E., VASQUEZ , E., DUVAL, B., LATREILLE, M. & COFFI-
ROURE, F., COLLETTA, B., DE TONI, B., LOUREIRO, D., PAS- NIERES, B. (1972) Paleogeographic and petroleum synthesis of
SALACQUA, H. & GOU, Y. (1997) Within-plate deformations in western Venezuela. Paris, Editions Technip, 59 pp.
the Maracaibo and East Zulia basins, western Venezuela. ZEVENBOOM, D. (1995) Dinoflagellate cysts from the Mediter-
Mar. Pet. Geol., 14, 139–163. ranean Late Oligocene and Miocene. Ph.D. Diss. Univ.
RULL, V. (2006) A high mountain pollen-altitude calibration set Utrecht, ISBN 90-393-1131-5: 221 pp.
for palaeoclimatic use in the tropical Andes. The Holocene, 16,
105–117.
SCHUBERT, C. (1982) Neotectonics of Bocono fault, Western Manuscript received 15 September 2014; In revised form
Venezuela. Tectonophysics, 85, 205–220. 1 April 2015; Manuscript accepted 26 July 2015.

© 2015 The Authors


26 Basin Research © 2015 John Wiley & Sons Ltd, European Association of Geoscientists & Engineers and International Association of Sedimentologists

You might also like