You are on page 1of 15

Marine and Petroleum Geology 19 (2002) 711–725

www.elsevier.com/locate/marpetgeo

Petroleum generation and migration in the Putumayo Basin, Colombia:


insights from an organic geochemistry and basin modeling study
in the foothills
Félix T.T. Gonçalvesa,*, Cesar A. Moraa, Fabio Córdobab, Edgar C. Kairuzb, Blanca N. Giraldoa
a
ECOPETROL, Instituto Colombiano del Petróleo (ICP), km 7 Autopista Piedecuesta, Bucaramanga AA 4185, Colombia
b
ECOPETROL, Vicepresidencia Adjunta de Exploración (AEX), Calle 37 # 8-43, 88 Piso, Bogota, Colombia
Received 2 January 2002; received in revised form 21 May 2002; accepted 23 May 2002

Abstract
This paper discusses the results of 1D modeling and geochemical analyses of rock, oil and gas samples performed in the foothills of the
Putumayo Basin, Colombia. A multivariate statistical assessment of biomarker data points to the existence of three oil groups, two of the
them similar to oil families reported in previous studies and related to the Villeta Formation source rocks, and a novel one tentatively
associated with the Caballos Formation source rocks. Geochemical data of oil and gas samples points to a complex reservoir filling history in
the foothills, with at least two migration pulses with distinct maturity levels. Vitrinite reflectance data together with modeling results shows
that the source rock is practically immature in the Putumayo foothills, even in the footwall of large thrust faults. The thermal and burial
reconstruction of a pseudo well located in the present cordillera area suggests that the source rocks have started petroleum generation and
expulsion during the Late Oligocene/Miocene. The oil and gas expelled during such a period have migrated over 80 km, reaching the present
foreland area. Subsequent tectonic burial and/or higher heat flows related to the Late Miocene/Pliocene Andean Paroxysm is thought to
account for a second pulse of migration of thermally evolved hydrocarbons limited to the foothills area. q 2002 Elsevier Science Ltd. All
rights reserved.
Keywords: Putumayo basin; Villeta formation; Biomarker; Basin modeling

1. Introduction the characterization of oil families and potential source


rocks (Córdoba et al., 1997; Kairuz, Córdoba, Moros,
The Putumayo Basin, with an approximate area of Calderón, & Buchelli, 2000; Ramón, 1996). In this work,
30,000 km2, is located in southern Colombia and represents new geochemical data of rock, oil and gas samples from
the northward extension of the Oriente Basin in Ecuador wells located in the Putumayo foothills are presented and
(Fig. 1). Both basins are part of the prolific sub-Andean compared with previous results. Geochemical methods
petroleum province that lies east of the Andean Cordillera. included Rock-Eval pyrolysis, total organic carbon and
Since the 1940s, the drilling of over 110 exploratory wells in visual kerogen analyses of cuttings and core samples, GC
the Putumayo Basin resulted in the discovery of 27 oil and GC –MS analyses of oils and organic extracts, and bulk
fields, with reserves of about 400 MMbbl of oil and 0.3TCF and isotopic analyses of natural gas samples. Two wells of
of gas. The largest petroleum accumulation is the Orito field the foothills and a pseudo-well located in the present
(Fig. 1), located in the eastern limit of the foothills area and Andean Cordillera area were submitted to thermal and
with oil reserves of approximately 250 MMbbl. kinetic 1D modeling. The integration of geochemical
Previous geochemical and petroleum system investi- analyses and basin modeling results with available geologic
gations performed in the Putumayo Basin have emphasized and geochemical data of the Putumayo and Oriente basins
provided basis for the discussion of petroleum generation
* Corresponding author. Fax: þ 57-7-644-5444. and migration history, and the timing of hydrocarbon charge
E-mail address: fgoncalv@ecopetrol.com.co (F.T.T. Gonçalves). and trap formation in the Putumayo Basin.
0264-8172/02/$ - see front matter q 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 2 6 4 - 8 1 7 2 ( 0 2 ) 0 0 0 3 4 - X
712 F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725

Fig. 1. Simplified map showing the tectonic framework of the Putumayo Basin and the location of the main oil fields of the Putumayo Basin (modified from
Córdoba et al., 1997). On the bottom, schematic W–E cross section displaying the structural style and stratigraphic succession in the foothills and foreland
areas.

2. Geologic background (Córdoba et al., 1997). The Cretaceous – Paleocene


sequence encompasses the coastal to shallow marine
2.1. Stratigraphic and tectonic summary sandstones and shales of the Caballos Formation, the marine
limestones, shales and sandstones of the Villeta Formation,
The sedimentary infill of the Putumayo Basin comprises and the shallow marine to continental siliciclastic sediments
three major sedimentary sequences separated by regional of the Rumiyaco Formation (Fig. 2). The Eocene – Miocene
unconformities (Fig. 2). The Triassic –Jurassic sequence sequence comprises the fluvial/alluvial conglomerates of the
(Motema Formation) is considered as the economic base- Pepino Formation, the lacustrine/fluvial shales, siltstones
ment and is composed of volcanic deposits (tuffs and lava) and sandstones of the Orteguaza Formation, and the
and shallow marine to continental siltstones and sandstones continental to transitional shales, conglomerates and
F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725 713

The final accretion of the Western Cordillera during the


Maastrichtian/Early Paleocene caused the inversion of
ancient normal faults (Cooper et al., 1995), uplift of the
Central Cordillera and a marked change of sedimentation
conditions from marine (Villeta Fm.) to continental
(Rumiyaco Fm.). Subsequently (Early Paleocene/Middle
Eocene), the increase in convergence rates between the
Nazca and South America plates continued inverting normal
faults, leading to the development of thrust and back-thrust
structures, erosion and deposition of the coarse siliciclastics
of the Pepino Formation (Casero et al., 1997; Córdoba et al.,
1997). The enhanced uplift of the Central Cordillera during
Middle Eocene to Middle Miocene times (Cooper et al.,
1995) induced a flexural subsidence and the deposition of
the Orito-Belén Group and the Ospina Formation. Finally,
the intense compression related to the Andean Orogeny
(Late Miocene to Recent) induced the uplift and inversion of
back-arc basin depocenters, which resulted in the formation
of the Eastern Andean Cordillera and the development of
the sub-Andean foreland basins (Casero et al., 1997;
Coletta, Hebrard, Letouzey, Werner, & Rudkiewicz, 1990;
Cooper et al., 1995).
As well as in other sub-Andean basins, two major
structural compartments can be recognized in the Putumayo
Basin: foothills and foreland (Fig. 1). The foreland (or
platform) compartment covers most of the basin area, being
characterized by a regional monocline that dips to the west,
locally affected by subtle anticlines, reverse and normal
faults. The foothills compartment encompasses two distinct
domains: to the east, (1) a thick-skin tectonics domain that
displays large NE –SW high angle thrust faults associated
with small back-thrusts; and to the west, (2) a folded-belt
domain that shows typical thin-skin tectonic features
involving the Upper Cretaceous/Tertiary section and large
Fig. 2. Generalized stratigraphic column of the Putumayo Basin (after thrust faults that put the Pre-Cretaceous sequence over the
Córdoba et al., 1997). Cenozoic sequence (Córdoba et al., 1997).
sandstones with minor gypsum and coal layers of the Orito-
Belén Group and the Ospina Formation. The Pliocene to 2.2. Petroleum systems
Quaternary coarse siliciclastic deposits of the Guamués and
The marine calcareous shales and marls of the Villeta
Caimán formations cover the older sequences. Igneous
Formation are considered the main petroleum source rocks
intrusions of inferred post-Oligocene age locally affect
of the Putumayo Basin (Córdoba et al., 1997; Giraldo &
the sedimentary section (Córdoba et al., 1997).
Ramón, 1997; Kairuz et al., 2000; Ramón, 1996; Rangel,
The highly complex tectonic-sedimentary evolution of the
Giraldo, & Córdoba, 1997). Such rocks display high total
Putumayo reflects the multiphase geological history of the organic carbon contents (TOC up to 10%), fair to good
northwest corner of South America (Casero, Salel, & Rossato, hydrocarbon source potential (S2 of about 10 mgHC/gRock)
1997; Cooper et al., 1995; Dengo & Covey, 1993). From Late and hydrogen indices (HI around 300– 400 mgHC/gTOC),
Triassic to Jurassic times, the siliciclastic and volcanic and dominance of amorphous organic matter. Coeval source
deposits of the Motema Formation filled grabens formed in a rocks are found within the Napo Formation to the south, in
back-arc setting (Casero et al., 1997). During Aptian to the Oriente Basin (Dashwood & Abbotts, 1990; Mello,
Maastrichtian times, the sedimentation of the Caballos and Koutsoukos, & Erazo 1995).
Villeta formations took place over a broad and relatively Two major oil families have been recognized the
shallow pericratonic depression that extended throughout Putumayo Basin (Córdoba et al., 1997; Ramón, 1996;
most of northwest South America. During this phase, relative Rangel et al., 1997). The so-called marine carbonate oil
sea-level changes and paleoceanographic events exerted a family, pooled within sandstone reservoirs of the Pepino and
primary control on sedimentation (Villamil & Pindell, 1998). upper Villeta (informal unit N) formations (Fig. 2), is
714 F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725

characterized by low to moderate gravity (, 308 API),


relatively high sulfur contents (around 1.5%), pristane/
phytane ratios around one, low Ts/Tm ratios, high relative
proportion of C35 homohopanes when compared to their C34
counterparts, and low diasteranes/regular steranes ratios.
The marine siliciclastic oil family, pooled within sandstone
reservoirs of the Caballos and lower Villeta (informal units
T and U) formations (Fig. 2), is marked by higher API
gravity values (up to 408), lower sulfur contents (, 0.75%),
pristane/phytane ratios ranging between 1 and 2, high Ts/Tm,
higher relative abundance of C34 homopanes, and higher
diasteranes/regular steranes ratios.
Biomarker-based oil– source rock correlations (Córdoba
et al., 1997; Ramón, 1996; Rangel et al., 1997) suggest that
the marine carbonate oils were sourced by Cenomanian –
Turonian shales and marls of A and M2 informal units
(upper part of the Villeta Formation, Fig. 2), while the
marine silicilastic oils are related to the Albian B and T
shales (lower part of the Villeta Formation). Based on such
correlations, Córdoba et al. (1997) and Kairuz et al. (2000)
Fig. 3. Geochemical profile of the Unicornio-1 well (for location see Fig. 1),
defined two distinct petroleum systems: Upper Villeta– showing the variation of total organic carbon content (TOC), hydrogen
Pepino and Lower Villeta–Caballos, with the later comprising index (HI) and vitrinite reflectance (%Ro) as a function of depth.
about 65% of the discovered reserves in the Putumayo Basin.
Visual kerogen analyses and pyrolysis data indicate that The Unicornio-1 well went through a large thrust fault
the organic-rich deposits of the Villeta and Napo formations (Fig. 3), finding the Villeta and Caballos formations
range from immature to early mature throughout the repeated both in the upthrown and downthrown blocks.
Putumayo and Oriente basins (Córdoba et al., 1997; Oil shows were found only in the hanging wall, within the
Dashwood & Abbotts, 1990; Mello et al., 1995). According Caballos Formation sandstones. The calcareous shales and
to previous basin modeling studies, the main phase of oil marls of the Villeta Formation show fair to high TOC
generation and expulsion was reached prior to the Andean contents (mostly from 1 to 3%), fair hydrocarbon source
Paroxysm in both basins (Córdoba et al., 1997; Dashwood & potential (S2 up to 9 mgHC/gRock), hydrogen indices from
Abbotts, 1990; Tegelaar, Zaugg, Hegre, & Rangel, 1995). 100 to 300 mgHC/gTOC, and a dominance of amorphous
organic matter (over 90%). In the calcareous shales of the
Caballos Formation, TOC contents locally reach up to 5%,
3. Organic geochemistry results but hydrogen indices are generally lower than 200 mgHC/
gTOC, which is in accordance with the higher proportion of
3.1. Source rock characterization woody organic matter (up to 70%). Vitrinite reflectance
measurements around 0.55 – 0.60% and Tmax values lower
Total organic carbon, Rock-Eval pyrolysis and visual than 435– 440 8C indicate that the Villeta Formation is
kerogen analyses (see Appendix A for analytical pro- immature to early mature both in the upthrown and
cedures) were performed on about 200 samples (cuttings downthrown blocks of the fault (Fig. 3). The tectonic burial
and core) from two wells of the foothills area: Orito-3 and related to the thrusting event has apparently not affected the
Unicornio-1 (Fig. 1). The results of bulk analyses of both thermal evolution of the footwall source rocks.
wells are in accordance with data from previous studies. In
the Orito-3 well (not shown in this paper), the Villeta 3.2. Oil characterization and oil– source rock correlation
Formation displays high TOC contents (up to 9%) and S2
values (up to 40 mgHC/gRock), hydrogen indices between Saturates fraction of 48 samples were submitted to gas
200 and 500 mgHC/gTOC), and a dominance (. 80%) of chromatography (GC) and gas chromatography – mass
amorphous organic matter. The Caballos Formation pre- spectrometry analyses (GC – MS; see Appendix A for
sents TOC contents ranging from 1 to 3%, fair to good procedures). Analyzed samples comprised 14 organic
hydrocarbon source potential (S2 up to 11 mgHC/gRock) extracts from organic-rich calcareous shales and marls of
and hydrogen indices between 120 and 350 mgHC/gTOC, the Villeta and Caballos formations in the Orito-3 and
reflecting the higher proportion of woody material (between Unicornio-1 wells, 21 crude oils produced from reservoirs
25 and 85%). Vitrinite reflectance and Tmax values reaches of the Pepino, Villeta and Caballos formations of several
0.6% and 435 8C in the Caballos Formation, which indicates wells of the Orito field, and 13 oils extracted (see Appendix
that the potential source rocks are immature in this well. A for procedures) from core samples of reservoir rocks of
F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725 715

the Caballos Formation in the Orito-3 and Unicornio-1 stage of thermal evolution (C29 steranes abb/aaa þ abb
wells. and aaa S/S þ R ratios lower than 0.65 and 0.48,
Besides a visual inspection of GC and GC – MS (m/z 191 respectively), oil – oil and oil – source rock correlations
and 217) fingerprints, a multivariate statistical approach was were likely unaffected by maturity. By using the multi-
used to differentiate and correlate the analyzed oils and variate statistical approach, three broad clusters are clearly
organic extracts. Primarily source-dependent biomarker identified in the resulting dendrogram (Fig. 4).
ratios, such as pristane/phytane, C24 tetracyclic/C26 tricyclic The cluster labeled A comprises lower Villeta Formation
terpanes, Ts/Tm, C35/C34 homohopanes, hopanes/steranes, (informal units B and T) organic extracts of the Orito-3 and
diasteranes/regular steranes and C30/C29 steranes (Peters & Unicornio-1 wells and a number of oils produced from the
Moldowan, 1993), were calculated and compared using the Caballos Formation in several wells of the Orito field
hierarchical cluster analysis technique available in the (Fig. 4). These samples are characterized by Pristane/
commercial software program DATADESK 6.0. Since source Phytane . 1, dominance of low molecular weight n-
rock samples are immature to early mature (around alkanes, high relative abundance of tricyclic terpanes, Ts/Tm
0.6%Ro) and oils have not exceeded the peak generation around 1, moderate abundance of C29 norhopane, C35/C34

Fig. 4. Hierarchical cluster analysis dendrogram showing the differentiation of three source-related families (A, B and C). Each sample is identified by the well
name, formation and type (organic extracts from source rocks, oils from production wells and oils extracted from reservoir cores).
716 F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725

homopanes ratios generally lower than 1, high diasteranes/ (Mello, Gaglianone, Brassel, & Maxwell, 1988; Mello et al.,
steranes ratios, dominance of C27 steranes over their C28 and 1995), while the molecular features of cluster B samples
C29 counterparts and moderate to high C30/C29 steranes ratio point to a more distal/deeper marine carbonate dysoxic–
(Fig. 5(a)). anoxic environment (Mello et al., 1988; Mello et al., 1995;
The cluster labeled as B embraces upper Villeta Palacas, Anders, & King, 1984; Zumberge, 1984). Geo-
Formation (informal units A and M2) organic extracts of chemical characteristics of clusters A and B oils display a
the Orito-3 and Unicornio-1 wells, oils produced from the good correlation with features reported by former authors
Upper Villeta and Pepino reservoirs in a number of wells of respectively for the so-called marine siliciclastic and marine
the Orito field and oils extracted from sandstone core carbonate oil families (Kairuz et al., 2000; Ramón, 1996;
samples of the Caballos Formation in the Orito-3 well Rangel et al., 1997). The distribution of the organic extracts
(Fig. 4). All such samples are characterized by Pristane/ between these clusters (Fig. 4) is also in accordance with
Phytane equal or lower than one, dominance of low such previous work, confirming that the marine siliciclastic
molecular weight n-alkanes, high relative abundance of oil family was probably generated by the lower Villeta
tricyclic terpanes, Ts/Tm , 1, moderate abundance of C29 Formation (informal units B and T), while the marine
norhopane, C35/C34 homopanes ratios equal or higher than carbonate oils are related to the upper part of the same
1, moderate to low diasteranes/steranes ratios, dominance of formation (informal units A and M2). Different from the
C27 steranes over their C28 and C29 counterparts and high previous studies however, marine carbonate oils were found
C30/C29 steranes ratio (Fig. 5(b)). also in the reservoirs of the Caballos Formation in the Orito-
The biomarker characteristics of cluster A samples are 3 well (Fig. 4), indicating that the upper and lower Villeta
indicative of a relatively proximal/shallow marine deposi- formations petroleum systems are not completely isolated as
tional environment with a significant siliciclastic input previously proposed by Ramón (1996). The confirmation of

Fig. 5. Gas chromatograms (GC) and m/z 191 and 217 mass chromatograms of representative oils of clusters A, B and C from Fig. 4. Peaks identification:
1 ¼ n-C15, 2 ¼ Pristane, 3 ¼ Phytane, 4 ¼ n-C25, 5 ¼ C23 tricyclic terpane, 6 ¼ C24 tetracyclic terpane, 7 ¼ C27 18a(H) Trisnorhopane (Ts), 8 ¼ C27 17a(H)
Trisnorhopane (Tm), X ¼ unknown terpane, 9 ¼ C29 17a(H), 21b(H) norhopane, 10 ¼ C30 17a(H) diahopane, 11 ¼ C30 17a(H), 21b(H) hopane, 12 ¼ C31
17a(H), 21b(H) homohopane 22S, 13 ¼ C27 ba diasterane 20S, 14 ¼ C27 abb sterane 20R þ C29 ba diasterane 20S, 15 ¼ C29 abb sterane 20R.
F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725 717

the lack of marine siliciclastic oils (derived from Lower The GC traces of the studied oil samples point to the
Villeta source rocks) in the Late Cretaceous and Tertiary existence of different histories of petroleum migration and
reservoirs of the Orito Field suggests that the Orito Fault alteration in the Putumayo foothills. Oils produced from the
behaved as a seal for petroleum migration. reservoirs of the Caballos Formation in the Orito field
The cluster labeled C encompasses organic extracts of display a normal distribution of n-alkanes (Fig. 6(a)) and a
organic-rich shales from the Caballos Formation in the lack of demethylated hopanes, while oils extracted from
Orito-3 and Unicornio-1 well and an oil extracted from a cores taken from the water zone of the same reservoir are
core sample of the Caballos Formation in the Unicornio-1 characterized by the depletion of n-paraffins and a
well (Fig. 4). Such samples are characterized by a strong pronounced unresolved complex mixture (‘hump’) that
dominance of pristane over phytane (Pr/Ph . 3), high rises above GC baseline (Fig. 6(b)). Such results indicate
proportion of high molecular weight n-alkanes (. n-C23) that the Orito petroleum accumulation was not biodegraded
with odd over even preference, low relative proportion of in the past (Peters & Moldowan, 1993) but alteration
tricyclic terpanes, Tm higher than Ts, presence of an processes are currently affecting the base of the oil column.
unknown terpane (X) that elutes just before the Tm, Conversely, the saturates fraction of oils extracted from core
significant abundance of C30 diahopane, high relative samples of the Caballos Formation (hanging wall) in the
proportion of C29 norhopane, very low C35/C34 homopanes Unicornio-1 well presents a depletion of high molecular
ratios, moderate to low diasteranes/steranes ratios, domin- weight n-alkanes, baseline hump and high proportion of
ance of C29 steranes over their C27 and C28 counterparts and pristane and phytane associated with a significant abun-
low C30/C29 steranes ratio (Fig. 5(c)). These features are dance of low weight n-alkanes (Fig. 6(c)). Such association
notably distinct from the characteristics displayed by points to a complex reservoir filling history, possibly
samples from previous clusters and points to a nearshore comprised by an early oil migration pulse, a subsequent
(transitional) marine environment with a strong input of biodegradation event that accounts for the removal of the n-
terrestrial plants (Abdullah, Murchinson, Jones, Telnaes, & paraffins, and a late migration pulse of lighter oil. The
Gjelberg, 1988; Isaksen, 1995; Philp & Gilbert, 1986). The absence of demethylated hopanes in such extracted oils
presence of organic extracts of the Caballos Formation from indicates that the paleobiodegradation event was not
both Unicornio and Orito areas together with the oil in intense.
cluster C (Fig. 4) suggests that the organic-rich shales of this
unit are the source rocks. Such results provide support to a 3.3. Natural gas characterization
proposition of a new speculative petroleum system
designated Caballos, which is subject to confirmation by Bulk and isotopic composition of seven gas samples from
further analyses. wells of the foothills and four gas samples from wells of the

Fig. 6. Whole oil gas chromatogram of an oil produced from the Caballos Formation reservoir in the Orito-35 at 1975 m depth (a), and gas chromatograms of
the saturate fraction of oils extracted from core samples of the Caballos Formation reservoir in the water zone of the Orito-3 at 2016 m depth (b) and in the
upthrow block of the Unicornio-1 at 1306 m depth (c). For peak identification see Fig. 5.
718 F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725

foreland area were used to assess gas generation and comprised by over mature gases enriched in isotopically
migration history in the Putumayo Basin. All analyzed heavy methane.
samples present high contents (up to 90 vol.%, Fig. 7) of
carbon dioxide with carbon isotopic ratios ranging from
2 1.3 to 2 7.1‰. Such high contents of relatively light CO2 4. Basin modeling
have been considered the result of the thermal breakdown of
carbonate rocks of the Villeta Formation by the heat effect To better understand the thermal evolution of the source
of the Tertiary igneous intrusions (Kairuz, 1993). Concern- rocks in the Putumayo foothills, the Orito-3 and Unicornio-1
ing the hydrocarbons, foothills samples are in general dryer, wells (Fig. 1) were submitted to one-dimensional thermal
displaying higher relative abundances of methane when and kinetic modeling using the software GENEX 3.3.0. The
compared gases from the foreland wells (Fig. 7). The structural complexity of the Unicornio-1 well required the
application of the interpretative approach proposed by use of the GenTect module.
Prinzhofer, Mello, Freitas, and Takaki (2000) and Prinzhofer Burial reconstruction took into account measured thick-
and Pernaton (1997) suggests that bulk composition and ness of formations drilled in the wells and erosion estimates
isotopic signature of hydrocarbon gases were primarily based on regional stratigraphic correlations (Casero et al.,
controlled by the degree of thermal evolution, without 1997; Córdoba et al., 1997). Absolute ages of depositional
significant influences of other phenomena such as segrega- and erosional events were defined using the chronostrati-
graphic framework of the basin (Fig. 2) and the geologic
tive migration (Fig. 8(a)) or biodegradation (Fig. 8(b)).
time scale proposed by Harland et al. (1989). Compaction
Since secondary phenomena appear to have not sub-
history was described through a porosity– depth law in
stantially affected gas composition, a model that describes
which porosity is a function of depth. An average porosity–
the isotopic evolution of methane and ethane generated by
depth curve was automatically calculated by the program
types I and II kerogens as a function of maturity (Whiticar,
for each formation by weighing the porosity law of each
1994) was used to assess the degree of thermal evolution of pure lithology (sandstone, shale, limestone, etc.) according
the analyzed samples. By applying such model, two groups to its relative proportion in the formation. Paleo-water
of gases with distinct characteristics may be recognized depths were not incorporated to the model due to the lack of
(Fig. 9). The foreland samples present a co-genetic behavior reliable estimates. This simplification probably will have
with isotopic ratios indicative of a moderate degree of little impact over the burial reconstructions, since the
thermal evolution (oil window). On the other hand, most of studied sedimentary succession was deposited under con-
the foothills samples display an association of isotopically tinental to relatively shallow marine environments (Córdoba
heavy methane indicative of a high degree of maturity (gas et al., 1997). The resulting geohistory plots (e.g. Orito-3
window) with a relatively light ethane that is more well, Fig. 10) shows two major subsidence phases (Late
compatible with the oil window. Such association points Cretaceous/Paleocene and Late Eocene/Miocene) and uplift
to the mixing of gases from at least two migration phases, events (Early/Middle Eocene and Late Miocene/Pliocene).
being one composed of mature gases and the other According to the burial reconstruction of the modeled wells,

Fig. 7. Compositional variation of the 11 analyzed gas samples taken the foothills and foreland areas of the Putumayo Basin. The graph displays the proportion
in volume of carbon dioxide (CO2), methane (C1) and hydrocarbon gases heavier than ethane (C2þ).
F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725 719

Fig. 10. Burial curves of the Orito-3 well, Putumayo foothills (see location
on Fig. 1).

geothermal gradients are about 20 – 21 8C/km. Thermal


conductivity of the sedimentary column was automatically
Fig. 8. Geochemical diagrams used to characterized the gases from the calculated by the program taking into account the
Putumayo Basin (after Prinzhofer & Pernaton, 1997; Prinzhofer et al., lithological composition, porosity decay and temperature
2000). The d 13C1 versus C2/C1 cross plot (a) shows a dominance of the
maturity trend over the segregative migration trend, while the C2/C3 versus of each formation. Due to the lack of measured data, thermal
C2/iC4 cross plot (b) indicates that the gases were not affected by conductivity values available in GENEX for pure litholo-
biodegradation. gies were used in the calculations. The radiogenic heat
contribution from the basement and the sediments was also
the base of the Villeta Formation reached its maximum included in the modeling by using default values of
depth (about 3000 m) during the Late Miocene. radiogenic heat production for pure lithologies available
The assessment of corrected bottom-hole temperatures in the modeling program. After a number of trial-and-error
(BHT) of over 100 wells indicates that mean geothermal simulations, the best fit between measured and calculated
gradients vary between 18 and 25 8C/km across the temperatures (Fig. 11) was obtained by using heat flow
Putumayo Basin. In the Orito-3 and Unicornio-1 wells, values of about 36 mW/m2 in the Orito-3 and 38 mW/m2 in
the Unicornio-1 well.
The highly complex tectonic evolution of the Putumayo
Basin hampers the reconstruction of its heat flow history.
Diverse paleoheat flow scenarios were simulated, such as
(1) a constant reduced heat flow through time, (2) a higher
heat flow during the Early Cretaceous rifting event, and (3) a
heat effect associated with the Tertiary igneous episode.
Nevertheless, despite the tectonic complexity, a constant
reduced heat flow equal to the present was the thermal
history scenario that best accounted for the measured
vitrinite reflectance values in both wells (Fig. 12).
Apparently, the thermal influence of mentioned tectonic
events on source rock maturity in the Putumayo Basin was
practically insignificant. Such results contrast with the ones
obtained for other Colombian basins (Mora & Santos Neto,
2000), where a strong thermal anomaly associated to the rift
phase is required to account for the present degree of
maturity of the source rocks.
Fig. 9. Ethane (d 13C2) versus methane (d 13C1) carbon isotopic ratios Kinetic determinations performed on a selected imma-
crossplot for foothills and foreland gases of the Putumayo Basin. The
diagram also displays the maturity trend with vitrinite reflectance
ture sample of the Villeta Formation source rocks revealed
equivalent values for co-genetic ethane and methane according to Whiticar activation energy distributions and frequency factors
(1994). representative of Type II kerogens (Fig. 13), with frequency
720 F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725

Fig. 11. Calibration of the present thermal regime in the Orito-3 (a) and
Unicornio-1 (b) wells, Putumayo foothills (see location on Fig. 1). Notice
the good correlation between measured (squares) and calculated tempera-
ture (thick dashed lines). Fig. 12. Calibration of thermal and maturity modeling in the Orito-3 (a) and
Unicornio-1 (b) wells, Putumayo foothills (see location on Fig. 1). Notice
the good correlation between measured (squares) and calculated vitrinite
factor of 1.0 £ 1014 s21 and principal activation energy of reflectance (thick dashed lines).
53 kcal/mol. Using these measured kinetic parameters and
the calibrated thermal model, kerogen transformation ratios approximately 80 km to the northwest of the Unicornio-1
calculated for the Villeta Formation in the studied wells are well (Fig. 1). The poor knowledge of the structural
lower than 10% (Fig. 14), confirming that the source rocks configuration in the Cordillera and the lack of seismic
are practically immature in the foothills. Besides, the small data hamper the definition of the present stratigraphic
differences found between transformation ratios calculated succession in the selected location. By assuming that the
above and below the thrust-fault in the Unicornio-1 well Cretaceous to Oligocene units succession was little
indicates that, at least in this area, the tectonic burial was not disturbed by pre-Miocene tectonic events, burial history of
enough to enhance substantially the conversion of kerogen such units was reconstructed based on the extrapolation of
into petroleum. sediment isopachs from the platform and foothills areas
Since mature source rocks have not been found in the (Córdoba et al., 1997) to the selected location. Lithologies,
platform and foothills, the hypothesis of the existence of an as well as their petrophysical and thermophysical
ancient oil-kitchen in the area of the present Cordillera was parameters were based on the composition of each
tested through the simulation of a pseudo-well located formation in the foothills. Likewise, the constant reduced
F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725 721

Fig. 13. Activation energies distribution and frequency factor of a representative sample of the Villeta Formation from a well located in the foreland area.

heat flow history calibrated with temperature and vitrinite the maturity of most of the oils and gases of the Putumayo
reflectance data in the Unicornio-1 well was used in the Basin, but it is lower than the maturity level of the light oils
pseudo-well. and over mature gases attributed to a later migration pulse in
According to the geohistory plot obtained for the pseudo- the foothills. Therefore, a higher heat flow and/or additional
well (Fig. 15), the base of the Villeta Formation might have burial would be required to lead the source rock to a higher
reached about 6000 m depth by the beginning of the stage of thermal evolution.
Miocene. Due to the lack of geologic data to constrain the
Miocene to Present burial and deformation histories, neither
additional stratigraphic/tectonic burial nor uplift events 5. Petroleum generation and migration: summary and
were included in the model during this period. Hence, discussion
results of burial and thermal modeling of the last 10 My are
disregarded. Modeling of kerogen conversion using the Pyrolysis and visual kerogen analyses performed in this
measured kinetic parameters indicates that Villeta For- study reasserts the results of previous works that found that
mation source rocks possibly started to generate petroleum the Villeta Formation source rocks have not reached the
during the Late Oligocene and that they attained a main phase of petroleum generation and expulsion in the
transformation ratio of over 80% and a maturity degree of foothills of the Putumayo Basin. Since the source rocks are
about 0.9%Ro before the Andean Paroxysm (Figs. 16 and
17). Such a degree of thermal evolution is consistent with

Fig. 14. Modeled evolution of kerogen transformation ratio of the middle Fig. 15. Burial curves of a pseudo-well located in the present cordillera area
part of the Villeta Formation in the Orito-3 and Unicornio-1 wells, of the Putumayo Basin (see location on Fig. 1). Burial history was
Putumayo foothills (see location on Fig. 1). reconstructed only until the Late Miocene (10 Ma).
722 F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725

Fig. 17. Modeled reconstruction of the maturity evolution of the middle part
of the Villeta Formation in terms of vitrinite reflectance (%Ro) until the
Late Miocene. Notice that before the Andean Paroxysm (about 10 Ma) the
source rocks have reached around 0.9%Ro, which is compatible with the
degree of thermal evolution of the oils and gases attributed to the early
migration pulse, but is not in accordance with the maturity level estimated
for the highly evolved gas and light oil probably related to a later pulse (see
text for explanation).

Oriente Basin, Dashwood and Abbotts (1990) report that the


productive structures show evidence of Cretaceous to
Fig. 16. Modeled evolution of kerogen transformation ratio of the middle Oligocene growth.
Villeta Formation until the Late Miocene (10 Ma) in a pseudo-well located Geochemical data of oil and gas samples point to a
in the present cordillera area of the Putumayo Basin (see location on Fig. 1). complex hydrocarbon charge history, with at least two
pulses of petroleum migration. The GC fingerprint of the oil
also immature in the foreland, the likely oil kitchen for the extracted from a core sample of the Unicornio-1 well
petroleum accumulations must have been located to the (Fig. 6(c)) points to a possible mixing of a biodegraded oil
west, in the deeper parts of the foothills and the present with a lighter and probably more mature oil from a later
cordillera. To the south, Dashwood and Abbotts (1990) pulse. Likewise, whole-oil chromatograms of petroleum
report that the Napo Formation (equivalent to the Villeta accumulations of the Oriente Basin foothills (e.g. Oglan
Formation) is also immature to early mature throughout the field; Dashwood & Abbotts, 1990) also point to the mixing
Oriente Basin, indicating that effective source areas are of biodegraded oil from an early pulse with a more mature
located to the west. oil. While hydrocarbon gases from the Putumayo foreland
Faults have probably exerted a minor role as conduits for share a moderate level of thermal evolution (oil window),
petroleum migration as evidenced by the lack of oils most of the gases from the foothills appear to comprise a
generated by the Lower Villeta Formation within the Pepino mixture of gases generated within the oil window with gases
Formation reservoirs in the Orito field, and by the relative from the (late) gas window. The integration of such
isolation between the two main petroleum systems through- geochemical information with the geological data suggests
out the basin (Córdoba et al., 1997; Ramón, 1996), despite that the less mature oils and gases are related to an early pulse
the local presence of oils generated by the Upper Villeta of migration that has probably preceded the main phase of the
Formation source rocks in the Caballos Formation reser- Andean Paroxysm, since it has charged traps located both in
voirs of the Orito-3 well reported in the present study. Since the foothills and the foreland. Conversely, the more mature
migration appears to have been dominantly horizontal, oils and gases, apparently restricted to the foothills, probably
hydrocarbon accumulations located in the foreland must correspond to a later pulse of migration that followed this
have formed before the development of the major thrust Andean Paroxysm, whose migration was limited by the
fault that separates the foothills from the foreland during the large thrust faults that disrupted the main carrier beds.
Andean Paroxysm (Late Miocene/Pliocene) and caused the The maturity modeling of the pseudo-well located in the
disruption of potential horizontal carrier beds (Fig. 1). present cordillera area (Fig. 1) indicates that the Villeta
Therefore, the existence of pre-Andean traps is required to source rocks might have reached the peak of oil generation
account for the formation of these accumulations. In the before the Andean Paroxysm due to an inferred westward
F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725 723

thickening of the sedimentary overburden (Figs. 15– 17). could also be invoked to account for the formation of the
Inasmuch as such a degree of thermal evolution does not highly mature hydrocarbons. However, since maturity and
account for the over mature gases and oils attributed to the temperature data point to the prevalence of relatively low
later migration pulse, an additional burial and/or a higher paleoheat flows throughout the entire Putumayo Basin, any
heat flow beyond those assumed for the pseudo-well would thermal event capable of providing additional heat flow
be required in the same region to enhance source rock should be restricted to the area of the present cordillera. The
maturity. The additional burial could have been provided by schematic diagram of Fig. 18 integrates and summarizes the
the tectonic overburden associated with the development of proposed model of hydrocarbon generation and migration in
the Andean Cordillera. Such a hypothesis implies the the Putumayo Basin.
existence of a pod of source rock bellow the present
cordillera, as well as a possible presence of highly evolved
hydrocarbon accumulations. Due to the final uplift and 6. Conclusions
erosion of the Andes, such a pod of source rock must be
presently inactive. Alternatively, a different thermal regime Results of bulk and visual kerogen analyses confirm that

Fig. 18. Diagram summarizing the proposed model of petroleum generation and migration in the Putumayo Basin. This simplified scheme is not to scale and
does not display the structural shortening.
724 F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725

the organic-rich deposits of the Villeta Formation are the WR-12 apparatus. Pyrolysis analyses were performed with a
main potential source rocks of the Putumayo Basin. Such Rock-Eval II apparatus following the method described by
deposits have not reached the oil window in the foothills, Espitalié, Deroo, and Marquis (1985). For the visual
even in the footwall of large thrust faults. A multivariate kerogen investigation, rock samples were demineralized
assessment of biomarker data of a set of oils and organic with HCl and HF. The isolated kerogen samples were
extracts points to the existence of three oil families: A, B mounted on strewn slides and plugs to be analyzed in
and C. The former two correspond to the marine siliciclastic transmitted (white light and fluorescence) and reflected
and marine carbonate oil families described in previous microscopy with Zeiss standard microscopes.
works, respectively generated by the source rocks of the
lower (B and T units) and upper (A and M2 units) Villeta. A.2. Extraction, GC and GC –MS analyses
The new so-called transitional marine (C) oil family is
provisionally related to the potential source rocks of the Powdered source rock samples were submitted to
Caballos Formation. Geochemical data from oil and gas Soxhlet extraction for 24 h using dichloromethane. Reser-
samples points to the existence of mixing of hydrocarbon voir rock samples were washed for 2 h also using the same
derived from at least two pulses of petroleum migration with solvent. Saturate, aromatic and NSO compounds fractions
distinct degrees of thermal maturity. of the recovered oils and organic extracts were eluted from a
The integration of geochemical data, 1D basin modeling silica-gel column using n-hexane, n-hexane þ toluene and
results and geologic evidence provide support to the toluene þ methanol, respectively. The saturates fractions
proposal of a tentative two-phase model of hydrocarbon were analyzed in a Hewlett –Packard 5890-A gas chro-
charge for the Putumayo Basin. During the first phase (Late matograph (splitless injector and 30 m SPB-TM-1 column),
Oligocene/Miocene), the Villeta Formation source rocks with hydrogen as carrier gas and column temperature
have reached the peak of oil generation. Oils and gases programmed from 120 to 310 8C at 6 8C/min. The same
expelled in this phase migrated over 80 km, filling pre- fractions were analyzed using a HP 5972A mass selective
Andean structures both in the foothills and foreland areas. detector coupled to a HP 5890A gas chromatograph (on-
During the second phase the later, concomitant and/or column injector and 30 m HP-5MS column) with helium as
subsequent to the Andean Paroxysm (Late Miocene/Plio- carrier gas and column temperature programmed from 70 to
cene), additional tectonic burial and/or heat flow gave rise to 170 8C at 20 8C/min and from 170 to 310 8C at 2 8C/min.
the high maturity hydrocarbons whose migration was The above geochemical analyses were carried out in the
restricted to the foothills by the major faults that separates laboratories of the Colombian Petroleum Institute (ICP).
this area from the foreland. Whole-oil gas chromatography analyses of oils from
production wells were performed by Biomarker Technol-
ogy, kinetic analyses of source rock samples by Corelab/
Acknowledgments Lab Instruments and gas analyses by DGSI and CENPES.

The authors extend their appreciation to Ecopetrol for


permitting the publication of the results of this study, to the
References
laboratory personnel from the Geochemistry Group of the
Colombian Petroleum Institute (ICP) for the analytical
Abdulah, W. H., Murchinson, D., Jones, J. M., Telnaes, N., & Gjelberg, J.
support, and to Antonio Rangel, Diego Garcı́a, Anthony (1987). Lower Carboniferous coal depositional environments on
Tankard and two anonymous reviewers for their thoughtful Spitsbergen, Svalbard. In L. Mattavelli, & L. Novelli (Eds.), Advances
reviews and comments. in organic geochemistry (pp. 953 –964).
Casero, P., Salel, J. F., & Rossato, A (1997). Multidisciplinary correlative
evidences for polyphase geological evolution of the foothills of the
Cordillera Oriental (Colombia). Proceedings of the VI Simposio
Bolivariano (pp. 100 –118), Cartagena, Colombia.
Appendix A. Analytical methods Coletta, B., Hebrard, F., Letouzey, J., Werner, P., & Rudkiewicz, J. L.
(1990). Tectonic style and crustal structure of the Eastern Cordillera
A.1. Total organic carbon, Rock-Eval pyrolysis and visual (Colombia) from a balanced cross-section. In J. Letouzey (Ed.),
Petroleum and tectonics in mobile belts (pp. 81–100). Paris: Editions
kerogen analyses
Technip.
Cooper, M. A., Addison, F. T., Alvarez, R., Coral, M., Graham, R. H.,
After being examined with a binocular microscope to Hayward, A. B., Howe, S., Martinez, J., Naar, J., Peñas, R., Pulham,
eliminate caved material, and drilling mud contaminants, A. J., & Taborda, A. (1995). Basin development and tectonic history of
rock samples were washed using fresh water, dried and wet- the Llanos Basin, Eastern Cordillera, and Middle Magdalena Valley,
Colombia. AAPG Bulletin, 79(10), 1421–1443.
sieved with a 10 mesh top sieve and a 16 mesh bottom sieve
Córdoba, F., Buchelli, F., Moros, J., Calderón, W., Guerrero, C., Kairuz,
and powdered. For the determination of organic carbon E. C., & Magoon, L (1997). Proyecto evaluación regional Cuenca del
contents (TOC), a small fraction of each sample was Putumayo—Definición de los sistemas petrolı́feros. Ecopetrol, Internal
decalcified with HCl (50%) and analyzed using a LECO Report, 140 p.
F.T.T. Gonçalves et al. / Marine and Petroleum Geology 19 (2002) 711–725 725

Dashwood, M. F., & Abbotts, I. L. (1990). Aspects of the petroleum ings of the Seventh Latin–American Congress on Organic Geochem-
geology of the Oriente Basin, Ecuador. J.Brooks, Geological society istry (pp. 389), Foz do Iguaçú, Brazil.
special publication no. 50, 89–117. Palacas, J. G., Anders, D. E., & King, J. D. (1984). South Florida Basin: A
Dengo, C. A., & Covey, M. C. (1993). Structure of the Eastern Cordillera of prime example of carbonate source rock of petroleum. J. G.Palacas,
Colombia: Implications for trap styles and regional tectonics. AAPG AAPG studies in geology, no. 18, 71–96.
Bulletin, 77, 1315–1337. Peters, K. E., & Moldowan, J. M. (1993). The biomarker guide:
Espitalié, J., Deroo, G., & Marquis, F. (1985). La pyrolise Rock-Eval et ses Interpreting molecular fossils in petroleum and ancient sediments.
applications. Revue de l’Institute Français du Pétrole, 40, 563–579. Englewood Cliffs, NJ: Prentice Hall, 363 p.
Giraldo, B. N., & Ramón, J. C (1997). Characterization and correlation of Philp, R. P., & Gilbert, T. D. (1986). Biomarker distributions in Australian
source rocks and crude oils, Putumayo Basin, Colombia. 18th oils predominantly derived from terrigenous source material. Organic
International Meeting on Organic Geochemistry, Abstracts Part I Geochemistry, 10, 73–84.
(pp. 377–378), Maastricht, The Netherlands. Prinzhofer, A., Mello, M. R., Freitas, L. C. S., & Takaki, T. (2000). New
Harland, W. B., Armstrong, R. L., Cox, A. V., Craig, L. E., Smith, A. G., & geochemical characterization of natural gas and its use in oil and gas
Smith, D. G. (1989). A geologic time scale. Cambridge: Cambridge evaluation. M. R.Mello, B. J.Katz, AAPG memoir 73, 107–120.
Prinzhofer, A., & Pernaton, E. (1997). Isotopically light methane in natural
University Press, 263 p.
gas: bacterial imprint of diffusive fractionation? Chemical Geology,
Isaksen, G. H. (1995). Organic geochemistry of paleodepositional
142, 193–200.
environments with a predominance of terrigenous higher-plant organic
Ramón, J. C. (1996). Oil geochemistry of the Putumayo Basin, Ciencia.
matter. A. Y. Huc, AAPG studies in geology, no. 40, 81–104.
Tecnologı́a y Futuro, 1, 25–34.
Kairuz, E. C (1993) Origen del CO2 en la Cuenca del Putumayo y su riesgo
Rangel, A., Giraldo, B. N., & Córdoba, F (1997). Geoquı́mica del petróleo
exploratorio asociado. Proceedings of the VI Congreso Colombiano de
de la Cuenca del Putumayo. Proceedings of the VI Simposio
Geologı́a (pp. 210–214), Medellı́n, Colombia. Bolivariano (p. 29), Cartagena, Colombia.
Kairuz, E. C., Córdoba, F., Moros, J., Calderón, W., & Buchelli, F (2000). Tegelaar, E. W., Zaugg, P., Hegre, J., & Rangel, A (1995). Petroleum
Sistemas petrolı́feros del Putumayo, Colombia. Proceedings of the VII systems of the foothills of the Southern Llanos and Putumayo Basins.
Simposio Bolivariano (pp. 525– 532), Bogota, Colombia. Proceedings of the 68 Congreso Colombiano del Petróleo (pp. 99 –106),
Mello, M. R., Gaglianone, P. C., Brassel, S. C., & Maxwell, J. R. (1988). Bogotá, Colombia.
Geochemical and biological marker assessment of depositional Villamil, T., & Pindell, J. L. (1998). Mesozoic paleogeographic evolution
environment using Brazilian offshore oils. Marine and Petroleum of Northern South America: Foundations for sequence stratigraphic
Geology, 5, 205 –223. studies in passive margin strata deposited during non-glacial times.
Mello, M. R., Koutsoukos, E. A. M., & Erazo, W. Z. (1995). The Napo J.Pindell, C.Drake, SEPM special publication, no. 58, 283 –318.
Formation, Oriente Basin, Ecuador: Hydrocarbon source potential and Whiticar, M. J. (1994). Correlation of natural gases with their sources.
paleoenvironmental assessment. In B. J. Katz (Ed.), Petroleum source L.Magoon, W. G.Dow, AAPG memoir 60, 261 –283.
rocks (pp. 167–181). Casebooks in earth sciences, New York: Springer. Zumberge, J. E. (1984). Source rocks of the La Luna Formation in the
Mora, C. & Santos Neto, E. V (2000). Heat flow versus overloading as Middle Magdalena Valley, Colombia. J. G.Palacas, AAPG studies in
driven forces of hydrocarbon expulsion in Colombian basins. Proceed- geology, no. 18, 127–134.

You might also like