Vous êtes sur la page 1sur 14

Geothermal Regime and Thermal History of the

Llanos Basin, Colombia1

Stefan Bachu,2 J. C. Ramon,3 M. E. Villegas,3 and J. R. Underschultz2

ABSTRACT Tertiary back-arc event in the west. Rapid cooling


since the last thermal event is possibly caused by
The Llanos basin is a siliciclastic foreland sub- subhorizontal subduction of cold oceanic litho-
Andean sedimentary basin located in Colombia spheric plate.
between the Cordillera Oriental and the Guyana
Precambrian shield. Data on bottom-hole tempera-
ture, lithology, porosity, and vitrinite reflectance INTRODUCTION
from all 318 wells drilled in the central and south-
ern parts of the basin were used to analyze its Two supergiant oil fields, Cano Limon and
geothermal regime and thermal history. Cusiana, have been discovered recently in the
Average geothermal gradients in the Llanos basin Llanos basin, a siliciclastic sub-Andean foreland
decrease generally with depth and westward basin located in Colombia between the Cordillera
toward the fold and thrust belt. The geothermal Oriental to the west and the Guyana Precambrian
regime is controlled by a moderate, generally west- shield to the east (Figure 1). These discoveries led
ward-decreasing basement heat flow, by deposi- to renewed interest in the basin and its hydrocar-
tional and compaction factors, and, in places, by bon potential. Given the link between temperature
advection by formation waters. Compaction leads and the various physical and chemical processes
to increased thermal conductivity with depth, leading to hydrocarbon generation, it is important
whereas westward downdip flow in deep sand- to understand the present geothermal regime and
stone formations may exert a cooling effect in the past thermal history of the basin.
central-western part of the basin. Vitrinite re- We used data collected by industry from 318
flectance variation with depth shows a major dis- wells drilled in the central and southern part of the
continuity at the pre-Cretaceous unconformity. Are- basin to evaluate the present-day geothermal regime
ally, vitrinite reflectance increases southwestward and the thermal history of the Llanos basin. These
in Paleozoic strata and northwestward in post-Paleo- data, of variable quality and uneven distribution
zoic strata. These patterns indicate that the ther- both areally and with depth, consist mainly of bot-
mal history of the basin probably includes three tom-hole temperature (BHT) and vitrinite re-
thermal events that led to peaks in oil genera- flectance (R o) measurements, and porosity logs.
tion: a Paleozoic event in the southwest, a failed Other types of data, like formation pressure and per-
Cretaceous rifting event in the west, and an early meability, were used in a study of the flow of forma-
tion waters in the Llanos basin (Villegas et al.,
1994); these results are being used here only in dis-
cussing the conductive or convective nature of the
©Copyright 1995. The American Association of Petroleum Geologists. All
rights reserved.
terrestrial heat flow in the basin. Because data quali-
1Manuscript received January 1, 1994; revised manuscript received May ty is generally quite variable, with errors occurring
31, 1994; final acceptance September 27, 1994.
2Alberta Geological Survey, Alberta Research Council, P.O. Box 8330,
during the various stages of measurement, collec-
Edmonton, Alberta, T6H 5X2, Canada. tion, analysis, and transfer to hard copy and elec-
3Instituto Colombiano del Petróleo, Guatiguara-Piedecuesta, Santander, tronic media, we analyzed and culled the data first
A.A. 4185 - Bucaramanga, Colombia. on an individual basis and then by whole-sample
This paper is the result of work performed by staff from Ecopetrol
(Empresa Colombiana de Petróleos), Instituto Colombiano del Petróleo (ICP), populations and categories. After checking, culling,
and the Alberta Research Council (ARC). Petro-Canada International and synthesizing the various data, we integrated all
Management Services (PCIMS), acting on behalf of the Canadian
International Development Agency (CIDA), provided funding support for
the available information into a model of the
performing the geothermal analysis at ARC headquarters in Edmonton, geothermal regime and history of the basin. This
Alberta. We wish to acknowledge and express our gratitude to our colleagues model can be used further in studying and numeri-
for their help, to ICP and PCIMS for their support, and to ICP, Ecopetrol, and
PCIMS for granting release of the material for publication. We would also like cally simulating the evolution of the Llanos basin
to thank David Deming for his constructive review of the manuscript. and in evaluating its source-rock maturation and

116 AAPG Bulletin, V. 79, No. 1 (January 1995), P. 116–129.


Bachu et al. 117

Figure 1—Location of the Llanos


basin in Colombia.

hydrocarbon potential. The analysis is divided into Precambrian shield; and to the south by the Ama-
two major parts. The main terrestrial heat sources, zon basin from which it is separated by the Serranía
thermal regime, and heat-flow mechanisms are ana- de la Macarena Mountains and the Vaupes base-
lyzed first using quantitative methods applied to ment arch. The low-lying grassland of the basin
measured BHT and rock-property data. Second, the ranges in altitude from more than 300 m in the
thermal history of the basin is inferred using qualita- west to less than 100 m in the east.
tive reasoning applied to heat-flow and Ro data dis- The basement underlying the Llanos basin con-
tributions. To provide a tectonic and geologic frame sists of very old (more than 2700 m.y.) remnants of
of reference, we present first a short description of high-grade metamorphic terranes that have been
basin setting, history, geology, and hydrogeology. remobilized in a large-scale thermal event at
approximately 2000 Ma. These were covered by
intensively metamorphosed sedimentary rocks in
BASIN HISTORY the 1000- to 1600-Ma age range that have been
intruded by acidic igneous rocks during the
The Llanos basin is toward the northern end of a Orinoco episode around 1300 Ma. Subsequently
long series of sub-Andean basins extending from metamorphosed upper Proterozoic sediments
Argentina to Venezuela, generally along the eastern (570–1000 Ma) underlie most of the Llanos basin,
side of the Andes Mountains. Most of the basin is which was initiated during the Paleozoic when a
located in Colombia, covering an area of approxi- large failed rift formed the Arauca graben in the
mately 200,000 km 2, with a small portion in the north. This graben was the site of extensive sedi-
northeast in Venezuela (Figure 1). The Llanos basin mentation that now makes up more than 8500 m of
is bounded to the west by the foothills of the Cambrian–Ordovician strata (McCollough, 1987).
Cordillera Oriental, the easternmost of the three An eastward-dipping subduction zone was estab-
distinct ranges of the Andes in Colombia; to the lished during the Late Triassic and Jurassic along
north by the Apure basin and Cordillera de Merida the western margin of South America, whose pres-
in Venezuela; to the east by the cratonic Guyana ent-day northwest corner was directly connected
118 Geothermal Regime, Llanos Basin, Colombia

Dominant
collision of a volcanic arc with the continental mar-
Period Epoch Formation gin, led to erosion of Cretaceous sediments and de-
Lithology
Quaternary Pleistocene position of Paleocene successions that prograded
Necesidad
Undifferentiated from the west (McCourt et al., 1984). Eocene ero-
Pliocene Sandstone
and sion caused regional truncation of Paleocene and
Guayabo Shale Cretaceous strata, creating a regional unconformity.
The last major depositional cycle in the Llanos
Miocene
Leon Shale basin began during the late Eocene. Uplift contin-
ued within the Cordillera Occidental and Central.
C1 Sandstone
Alternating transgression and regression within the
Tertiary C2 Shale
C foreland basin resulted in the deposition of inter-
a
r
C3 Sandstone bedded marine shales and marine to nonmarine
b C4 Shale sandstones. The Cordillera Oriental uplift during
Oligocene o
n
C5 Sandstone the middle and late Miocene Andean orogeny seg-
e C6 Shale mented the large foreland basin that existed since
r
a C7 Sandstone the Paleocene into the present-day Llanos basin in
C8 Shale the east and the Magdalena basin in the west (Fig-
Eocene Mirador Sandstone
ure 1). The balance of this depositional cycle is
Los Cuervos Shale
comprised of thick Pliocene and Pleistocene suc-
Paleocene Barco Sandstone
cessions containing material eroded from the
Guadalupe Sandstone
Cordillera Oriental, and is almost entirely preserved
in the Llanos basin. No Holocene erosion has taken
Cretaceous Late Gachetá Shale place in the basin except for a narrow area about
Une Sandstone 10–15 km wide along the fold and thrust belt,
Jurassic
where geological and sedimentological data sug-
gest that up to 1200 m of sediments have been
Triassic
removed (Peñalosa and Ramirez, 1993). Figure 3
shows a northwest-southeast–dip cross section
Paleozoic Devonian Undifferentiated through the center of the Llanos basin.
Ordovician Siliciclastics
Hydrogeologically, the Llanos basin is comprised
Unconformity Eroded or not deposited of a succession of sandstone aquifers and shaly
aquitards (Villegas et al., 1994). The porosity of the
Figure 2—Stratigraphic nomenclature and dominant sandstone formations is high (22% on average) and
lithology of the Llanos basin. decreases generally with increasing depth (Figure
4) most probably as a result of compaction. Hori-
zontal permeability, measured in core, is also high,
with North America and Africa (Pindell et al., 1988; on the order of 10 –13 to 10–12m2 on average. The
Dengo and Covey, 1993). East of the resulting mag- vertical anisotropy of sandstone permeability is 0.7.
matic arc, back-arc extension provided accommo- The flow of formation waters in the Llanos basin is
dation for Triassic and Jurassic red beds found in driven generally by topography from high-elevation
the Arauca graben (Maze, 1984; McCollough, areas in the southwest at the Serranía de la Macare-
1987). The Permian–Carboniferous interval seems na Mountains and in the west along the foothills, to
to represent a hiatus, although information regard- low-elevation areas along the Arauca and Orinoco
ing this period is largely absent (McCullough, rivers in the northeast (Villegas et al., 1994). In the
1987). central-western part of the basin, the flow of for-
A major unconformity separates Cretaceous mation waters in deep pre-Carbonera units is north-
from older strata. An early Cretaceous extension westward downdip, driven by pore space rebound
created a graben where the Cordillera Oriental is as a result of erosional unloading along the fold and
today; this graben was filled with up to 3 km of thrust belt (Villegas et al., 1994). Lateral hydraulic
Lower Cretaceous marine sediments (Dengo and gradients range between 0.6 and 10 m/km. The
Covey, 1993). Deposition during the Late Creta- main directions of formation water flow are illus-
ceous was characterized by alternating transgres- trated diagrammatically in the inset of Figure 5.
sive and regressive cycles of westward-prograding,
craton-derived clastic sedimentation and open-
marine shales (Figure 2). The end of the Cretaceous GEOTHERMAL REGIME
marks the transition from the passive margin to the
convergent margin in northwest South America. The geothermal regime of any sedimentary basin
The first in a series of orogenic movements, the is determined by climate, rock thermal properties,
Bachu et al. 119

A A' Figure 3—Structural dip cross sec-


300 Ground Surface
tion, Llanos basin.
0
Necesidad and Guayabo

1000 C1
Elevation above mean sea level (m)

Leon C2
C3
C4
C6 C5
C7
C8
2000 Mirador
Guadalupe
Gachetá
Une Venezuela
Precambrian Colombia
3000 Paleozoic

a
er
ill
rd
Co
n
A

er
st
Ea
Bogota
4000

A' Basin
of
Los Cuervos e

g
Guyana Shield

Ed
Barco
5000
0 25 50 km

Scale

and the strength and interaction of various heat The multiannual air temperature ranges between
sources and transfer mechanisms by which terres- 24°C at altitudes of 350 m in the foothills of the
trial heat is transported to the surface. In the Cordillera Oriental and 28°C at altitudes of 90 m in
absence of volcanic and tectonic activity, the main the lowlands close to the Guyana Precambrian
sources of heat in the undeformed Llanos basin are shield. The BHT data collected by industry tend to
the heat that flows upward from the mantle and be incomplete and contain errors, and have to be
the heat that is generated internally by the decay of screened with care (Chapman et al., 1984). Out-of-
radioactive elements. Other heat sources, such as range temperature records, such as 250°C at 300 m
phase changes and chemical reactions, are minor depth, or values less than the ground temperature,
(Rybach, 1981). The main mechanisms for heat were rejected from the database. Incomplete or
transfer in a sedimentary basin are convection by erroneous measurements were also culled (e.g.,
moving fluids and conduction. The importance of multiple BHT readings with no measurement time
convective vs. conductive heat transport in the given, or recording the same value, or with non-
basin depends mainly on rock permeability, and monotonically increasing values). We estimated the
can be assessed based on field data using dimen- formation temperature, Tf, using the Horner meth-
sional analysis (Bachu, 1985, 1988; van der Kamp od for the measurements with multiple, monotoni-
and Bachu, 1989). cally increasing BHT readings (Bullard, 1947; Chap-
man et al., 1984; Deming, 1989). Because the
circulation time, tc, was not recorded in any mea-
Data Processing surement, we used the value tc = 7 h obtained as an
average of circulation times recorded in the Alberta
We assessed the geothermal regime in the Llanos basin (Bachu and Burwash, 1991). This value was
basin using data collected by industry on BHT, found to lead to better linear fits in the Horner anal-
lithologies, and rock properties. First, we estimated ysis than the value tc = 4 h used for the Uinta basin
the ground surface temperature, Ts, from the multi- (Chapman et al., 1984). For single BHT measure-
annual average air temperature, because the former ments we applied a statistical correction with
is generally higher by only 2–3°C than the latter in depth to estimate the formation temperature
areas with a tropical or equatorial climate like the according to a modified version (Bachu and Bur-
Llanos basin. We obtained the areal distribution of wash, 1991) of the method suggested initially by
multiannual average air temperatures in the basin AAPG (1976) and improved subsequently by Wil-
from climatic data recorded at 36 weather stations lett and Chapman (1986). The best regression
across the basin by the Instituto Colombiano de (coefficient of correlation R2 = 0.8) was found to
Hidrología, Meteorología y Adecuatión de Tierras. be between the relative temperature correction
120 Geothermal Regime, Llanos Basin, Colombia

Porosity φ (%) culling operation we assumed also that the main


mechanism for heat transfer in the basin is conduc-
0 5 10 15 20 25 30 35 tion. For a sedimentary succession comprised
0 entirely of sandstone with high thermal conductivi-
ty (7 W/m°C) and low porosity (5%), the minimum
φ = 30.3 exp(-0.17815z) possible geothermal gradient corresponding to low
heat flow is 6.5°C/km. For a sedimentary succes-
sion comprised entirely of shales with low thermal
1 conductivity (1.2 W/m°C) and high porosity (30%),
the maximum possible geothermal gradient corre-
sponding to high heat flow is 102°C/km. In reality,
the sedimentary succession comprises both shales
Depth z (km)

2 and sandstones with varying properties. Thus, for a


siliciclastic basin with low, flat topography like the
Llanos, and in the absence of unknown very local
anomalous processes, the geothermal gradient val-
ues must lie within these extreme bounds. A few
3 singular BHT values leading to average geothermal
gradients outside this range were rejected. Finally,
we used same-well profiles of temperature varia-
tion with depth and areal temperature distributions
by formation to identify apparent thermal anoma-
4 lies that were individually checked. At the end of
this screening process, the resulting final data set
consisted of 802 BHT measurements in 215 wells.
No temperature corrections were applied for either
5 topographic or uplift/erosion effects because these
Well-scale value Regression line are negligible or nonexistent at the depths involved
(Henry and Pollack, 1988) and for the particular
Figure 4—Variation with depth of well-scale porosity in conditions of the Llanos basin (flat relief and almost
sandstone formations, Llanos basin. total lack of Holocene erosion).
We used data from 15 wells in the basin with
both lithological logs and vertical temperature pro-
∆T/T (in percent) and depth z (in kilometers), of files based on BHT to estimate the heat flow at the
the form top of the Paleozoic succession. Using the thermal-
resistance method we calculated the heat flow, taking
∆ T T = 16.9807 z − 9.0420 z 2 + 1.8868 z 3 − 0.1392 z 4 into account the dependence of thermal conductiv-
ity on temperature based on porosity, sand-shale
Based on formation and ground surface tempera- content, water and rock thermal conductivity, and
tures, we calculated the average geothermal gradi- geothermal gradients, similar to heat-flow calcula-
ent G between the ground surface and the BHT tions for the Alberta basin (Bachu, 1993). In the
measurement depth z according to absence of thermal-conductivity measurements for
rocks in the Llanos basin, we used the following
(
G = T f − Ts ) z values from the literature for matrix conductivities
from the Altiplano (in Bolivia) and Oriente (in
Because original information was not available Peru) sub-Andean sedimentary basins: 1.8 W/m°C
for checking some BHT data leading to very for shale and 4.0 W/m°C for sandstone (Henry and
anomalous average geothermal gradients, such as Pollack, 1988). These basins have a similar litholo-
3.8°C/km for very large depths or 270°C/km for gy, i.e., they are overwhelmingly marine and non-
shallow depths, we used a range-limiting procedure marine clastics, are characterized by both low and
to further discriminate and cull erroneous BHT high geothermal gradients (17 to 74°C/km) (Henry
data. We assumed a priori that the basement heat and Pollack, 1988), and have a similar evolution to
flow in the Llanos basin may range between 40 and that of the Llanos basin. The assumed matrix ther-
100 mW/m2, based on typical range of variation for mal conductivity leads to an average conductivity
heat flow in cratonic shield areas and mountain- of the sedimentary column on the order of 1.6
building regions in South America (Pollack and W/m°C. The obtained heat-f low values range
Chapman, 1977; Vitorello et al., 1980; Henry and between 29 and 63 mW/m2. We estimated the con-
Pollack, 1988; Nyblade and Pollack, 1993). For this tribution to heat f low from heat generated by
Bachu et al. 121

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000 Gauss


coordinates Figure 5—Areal distribu-
Venezuela tion of the average
Colombia geothermal gradient,
30
N Llanos basin (contour
interval = 5°C/km). Dots

25
20
1,200,000
1,200,000
represent data points. Inset
35
shows main groundwater
30

40
30 flow directions in Creta-
35
40 ceous–Eocene aquifers
1,100,000
35 45
1,100,000
(solid arrows) and in post-
ra

20
dille Eocene aquifers (open
or 35
n
C 30 50
arrows) (after Villegas et
er 35
as
t al., 1994).

30
E

40
25
20

30
1,000,000 Bogota 35 1,000,000

45
25

35 40
25

50
40
30

900,000 900,000

55
45
45
50

40
in
Bas
35

of Guyana Shield
25

e
45
50

g
Ed Scale
50

40
60
55 35 0 50 100 km
800,000 800,000
1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000

radioactive decay in the post-Paleozoic sedimentary estimates is not necessarily increased significantly.
rocks using the following values from the literature As pointed out by Bodell and Chapman (1982),
(Rybach, 1988) for radioactive heat production: 1.8 compiled values can be used as a last resort in the
µW/m3 for shale and 0.7 µW/m3 for sandstone. The absence of any measurements. Thus, although the
calculated heat f low generated by radioactive uncertainty in the present heat-flow estimates for
decay in 1000 m of mainly shale sediments is in the the Llanos basin is relatively high, these estimates
range of 1 mW/m2, showing that the contribution are still useful for regional-scale analysis by indicat-
of the sedimentary column is basically negligible in ing possible trends and order of magnitude. The
comparison to the basal heat flow. level of confidence in these heat-flow estimates
The calculated heat-flow values are estimates based on complete logs of porosity and sand-shale
only because of large uncertainties introduced by fraction is on the same order as estimates based on
the absence of direct measurements of shale and several chip or plug thermal conductivity measure-
sandstone matrix thermal conductivity. The values ments out of hundreds of meters of stratigraphic
used in calculations for all other variables (tempera- column. The confidence in our estimates is indi-
ture, porosity, sand-shale fraction) have a higher rectly enhanced by the fact that the calculated heat-
degree of confidence as a result of measurement flow values are in the range of measured values for
and data checking and culling. Given the range of the Brazilian Precambrian shield (Vitorello et al.,
variation of sandstone and shale thermal conductiv- 1980) and the sub-Andean Oriente basin in Peru
ity reported in the literature, we assessed the (Henry and Pollack, 1988) situated southwest of
uncertainty associated with the present heat-flow the Llanos basin. Future systematic thermal con-
estimates for the Llanos basin to be on the order of ductivity measurements should increase the degree
20–30%. The confidence in heat-flow determina- of confidence in heat-flow calculations.
tions can be significantly increased by measuring a
very large number of rock samples to characterize
adequately the stratigraphic column (Brigaud et al., Geothermal Gradients and Heat Flow
1990); however, in most cases the structural and
lithological complexity of a region is described by a Formation temperatures obtained from BHT in
limited number of samples obtained in holes of the post-Paleozoic succession in the Llanos basin
opportunity, each sample representing a few cen- range between 53 and 143°C. The temperature dis-
timeters depth interval (Bodell and Chapman, tribution in each stratigraphic unit exhibits an
1982). Thus, even if thermal conductivity measure- expected westward increase corresponding to stra-
ments are available, the confidence in heat-flow ta dipping toward the fold and thrust belt because
122 Geothermal Regime, Llanos Basin, Colombia

of the strong dependence of temperature on depth. G (o C/km)


The average geothermal gradient eliminates from
analysis the depth dependence of the thermal field, 0 10 20 30 40 50 60 70 80
allowing the possible identification of other pro- 0
cesses and phenomena. The areal distribution of the
average geothermal gradient (Figure 5) shows a defi-
nite trend of northwestward decrease from highs in
the 50°C/km range in the east near the Guyana Pre- 1000
cambrian shield to lows in the 20°C/km range in the
west–northwest along the fold and thrust belt. Sev-
eral local highs are superimposed over this general
trend, particularly in the south, indicating high
basement heat flow induced by a combination of 2000
basement highs present in the area (Muñoz, 1991;
Villegas et al., 1994) and basement granitic intru-
sions. Considering the dip of the strata from the
3000

Depth (m)
shield to the fold and thrust belt and, accordingly,
increased depth of BHT measurements, the north-
westward decrease in average geothermal gradients
suggests also a decrease with depth of these gradi-
ents. Indeed, the calculated values of the average 4000
geothermal gradients decrease with depth from
75°C/km between surface and 300 m depth to
18°C/km and less for large depths (Figure 6). The
large scatter of values at shallow depths is probably 5000
because the values correspond to BHT measure-
ments taken in wells distributed over the entire
basin, covering areas of high, moderate, and low
geothermal gradients in the southeast, center, and 6000
west–northwest, respectively (Figure 5). The lack of
scatter for great depths (Figure 6) is because these
depths are encountered only in a narrow area in the
west–northwest along the fold and thrust belt, 7000
which is characterized by low geothermal gradients
(Figure 5). Although the scatter at shallow depths Figure 6—Variation with depth of the average geother-
(Figure 6) could be partly attributed to noise in BHT mal gradient in the Llanos basin.
data, this is less probable in view of the definite
areal trend shown in Figure 5.
This pattern of decrease in average geothermal
gradients both northwestward and with depth (Fig- resulting in higher geothermal gradients for the
ures 5, 6) is most probably the result of the com- same heat flow. Also, porosity in the Llanos basin
bined effect of variations in thermal conductivity generally decreases with depth from 30% near the
with depth, advection by downdip flow of forma- surface to 13% at 5000 m depth due to compaction
tion water, and variable basement heat flow. Each (Figure 4). This means that the thermal conductivi-
of these factors is discussed individually in the fol- ty of the water-saturated rocks at great depths is
lowing paragraphs; however, establishing their rela- higher than that of rocks at shallower depths
tive importance requires sensitivity analysis based because of reduced water content, whose thermal
on quantitative modeling, which is beyond the conductivity is approximately 0.6 W/m°C. This
scope of this study. Qualitative reasoning and effect is offset in part by the decrease in thermal
dimensional analysis are used to discriminate conductivity with increasing temperature, hence
among these three factors. depth, for both water and rocks (Chapman et al.,
The variation of thermal conductivity with depth 1984; Deming and Chapman, 1988). Higher ther-
is due to both deposition and compaction. Deposi- mal conductivity for the deeper strata means slight-
tionally, the proportion of shales is higher toward ly higher effective thermal conductivity for the
the top of the sedimentary succession. Because, entire sedimentary column in the western part of
except for coals, shales have the lowest thermal the basin, resulting in lower average geothermal
conductivity of all sedimentary rocks, the effective gradients for the same heat f low. Thermal resis-
thermal conductivity of the upper strata is lower, tance estimates based on stratigraphic thickness,
Bachu et al. 123

lithology, porosity, and temperature variations basin. Thus, it is possible that lateral formation
show that the depositional and compaction effects water flow in Cretaceous–Eocene aquifers induces
may account for a variation in average geothermal a relative “cooling” effect in the central-western
gradients by a factor of 1.5. They cannot account area of the basin. Numerical modeling is needed to
for the four-fold range of variation in integral further quantify the effect of advective heat trans-
geothermal gradients and particularly do not port by formation waters.
explain the whole-range variation at shallow depths Even if advective hydrodynamic effects are as
(Figures 5, 6). important in places as heat conduction, they still
The flow of formation waters may also play a cannot account for the entire range and basin-wide
role in determining the geothermal regime in a sed- pattern of variation in geothermal gradients for sev-
imentary basin. At the basin scale, the flow of for- eral reasons. First, the flow of formation water is to
mation waters in the post-Eocene aquifers in the the west–northwest only in the central-western
Llanos basin (C1, C3, C5 and C7, Figures 2 and 3) is part of the basin in the Une and Mirador-to-
from the southwest to the northeast (inset of Fig- Guadalupe aquifers; in the post-Eocene C1, C3, C5,
ure 5) (Villegas et al., 1994). This direction is nor- and C7 aquifers, the flow is mainly from southwest
mal to the northwestward decrease in average to northeast (Villegas et al., 1994). Second, some
geothermal gradients (Figure 5). In the Creta- high heat-flow areas in the south and southeast
ceous–Eocene aquifers (Une and Guadalupe-to- coincide with local groundwater recharge. If advec-
Mirador, Figures 2 and 3), the flow of formation tion effects would be significant, the heat f low
waters is northeastward in the southwest and should be lower than in discharge areas, as in other
northeast (Villegas et al., 1994), again normal to the basins (e.g., Alaska North Slope) (Deming et al.,
general direction of geothermal gradient decreases. 1992), which is the opposite to the observed pat-
In the central-western part of the basin, the flow of tern. Third, hydrodynamic effects are absent in the
formation waters in these aquifers is downdip top Miocene and later shaly formations that blanket
northwestward (inset of Figure 5) (Villegas et al., the basin and reach a total thickness of more than
1994), driven by disequilibrium pressures caused 4000 m at the fold and thrust belt. The greatest
by erosional pore-space rebound. In the south– variation in geothermal gradients is found in these
southeast there is some recharge flow oriented formations (Figure 6). Thus, the most probable
northward, driven by local topographic highs. The cause for the observed geothermal pattern must be
general mismatch between the directions of forma- a generally westward decrease in basement heat
tion water f low and the variation of average flow. Unfortunately, only six wells reach the base-
geothermal gradients indicates that, on a basin- ment, and none of them has information needed
wide scale, advection by fluid flow is not an impor- for heat-flow calculations. The heat flow at the top
tant factor. Only in the central-western part of the of the Paleozoic was estimated in 15 wells in the
basin the downdip lateral flow of formation waters undeformed basin based on lithology and porosity
in the Une and Mirador-to-Guadalupe aquifers may logs and temperature profiles. The heat flow is gen-
exert a cooling effect. The importance of advective erally low, ranging between 29 and 63 mW/m 2
cooling in this region can be assessed by calculat- (average 42 mW/m2), in the range of measured val-
ing the dimensionless geothermal Peclet number, ues for the Brazilian Precambrian shield (Vittorello
Pe* (Bachu, 1988; van der Kamp and Bachu, 1989). et al., 1980) and the sub-Andean Oriente basin in
The geothermal Pe* takes into account the thermal Peru situated southwest of the Llanos basin (Henry
and hydraulic properties of the water-saturated and Pollack, 1988). This range shows that, basically,
rocks (porosity, permeability, thermal conductivity the heat flow in the Llanos basin reflects the transi-
and capacity, density, and viscosity) and the geome- tion from the tectonically stable Guyana Precambri-
try and thermal and hydraulic gradients of the flow an shield in the east to the Andean subduction zone
system. The geothermal Peclet number for the in the west. Moreover, the distribution of heat flow
deep aquifers in the central-western part of the at the top of Paleozoic strata (Figure 7) shows high-
Llanos basin was estimated at a regional scale in the er values along the eastern and southern flanks of
same way as for aquifers in the Alberta basin the basin than in the central and western parts. A
(Bachu and Cao, 1992). Because of uncertainties in decrease in heat flow toward the west–northwest
estimating various parameters, particularly could explain the west–northwestward decrease in
hydraulic and thermal conductivities and gradients geothermal gradients. This heat-flow trend must be
along bedding in sloping aquifers, the geothermal real because variations in thermal conductivity and
Peclet number calculations indicate orders of mag- hydrodynamic effects alone cannot explain the
nitude only. The obtained range of variation for Pe* observed pattern and range of geothermal gradient
between 10–2 and 101 indicates that heat transport variations.
by formation water cannot be discounted as negli- In summary, the present structure of the
gible in certain aquifers and areas in the Llanos geothermal field in the Llanos basin is probably the
124 Geothermal Regime, Llanos Basin, Colombia

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000 Gauss


coordinates Figure 7—Heat flow
Venezuela (mW/m2) at the top of the
Colombia Paleozoic, Llanos basin.
N

1,200,000 1,200,000

1,100,000
ra 1,100,000
lle
di
or
C 35.4
n 29.2
er 26.8 y
st 37.2
dar
Ea 51.9
Bou
n
35.4 38.1 zoic
l eo
1,000,000 Bogota 38.6
Pa 1,000,000
38.8
48.5

33.6
41.4
62.7
55.0

900,000 900,000

54.4
Ba s in
of Guyana Shield
e
g
Ed Scale

0 50 100 km
800,000 800,000

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000

combined result of the following factors, enumerat- Geological data show that, although pre-Creta-
ed in the most likely order of importance: (1) vari- ceous erosion was significant, very little erosion
ability in basement heat flow; (2) higher thermal has taken place since then (McCollough, 1987;
conductivity of rocks in the deeper strata due to Peñalosa and Ramirez, 1993). Continuous com-
decreased porosity caused by compaction; (3) paction (porosity) and Ro trends in the post-Paleo-
downdip west–northwestward flow of formation zoic succession indicate that an Eocene erosional
water in Cretaceous aquifers in the central-western event, which removed parts of Paleocene and some
part of the basin; and (4) deposition of thick, low- Cretaceous strata, was relatively minor. Recent ero-
thermal-conductivity shale at the top of the sedi- sion took place in only a narrow area (10–15 km
mentary succession. The relative importance of wide) along the fold and thrust belt, where it
these factors could be better estimated by perform- removed the equivalent of up to 1200 m of sedi-
ing thermal conductivity measurements and ments (Peñalosa and Ramirez, 1993). Thus, it fol-
numerical simulations of fluid and heat flow in the lows that the post-Paleozoic strata in the western,
basin. eroded part of the undeformed Llanos basin have
been uplifted by approximately 1000 m following
maximum burial. The post-Paleozoic rocks in the
THERMAL HISTORY eastern, uneroded part of the basin are currently at
their maximum burial depth. Distributions of Ro,
Knowing the thermal history of a sedimentary both areally by formation and with depth, indicate
basin is important for assessing the potential and that, because they are at or close to their maximum
timing of hydrocarbon generation. We used vitri- burial depth, only the rocks found in the lower-
nite reflectance (Ro) data to determine, in a qualita- most Cretaceous Une and Gachetá formations at
tive way, the thermal history of the Llanos basin. depths greater than 4000 m in the west-northwest-
We checked and culled for erroneous values some ern deep part of the basin are currently in the oil
570 R o data obtained from laboratory measure- window (Ro >0.6%). To illustrate this point, Figure
ments, first individually (e.g., R o values of <0.2% 8 shows the Ro distribution in the Une Formation,
were rejected) and then using same-well profiles of which unconformably overlies the Paleozoic strata.
Ro variation with depth, based on the fact that Ro The inference, based on R o, that these rocks are
increases generally with depth. At the end of this currently in the stage of oil generation is corrobo-
process, we used the remaining 287 Ro data to eval- rated by the depth of the 120°C isotherm (Figure
uate the burial and thermal history of the basin 9), the generally accepted temperature at which
starting from the present and going progressively intense oil generation starts (Tissot and Welte,
back in time. 1984). The depth of the isotherm was calculated
Bachu et al. 125

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000 Gauss


coordinates Figure 8—Distribution of
Venezuela vitrinite reflectance in the
Colombia Cretaceous Une Formation,
N Llanos basin (contour
interval = 0.05%). Dots rep-
1,200,000 1,200,000
resent data points.

1,100,000 1,100,000
ra
d ille
or
C
n
er
st 0.4

7
Ea
0.
ary

0.6
und
Bo

5
Bogota

0.
1,000,000 1,000,000
on
ati
0.5

rm
Fo
0.6
0.5

900,000 900,000

s in
o f Ba Guyana Shield
e

g
Ed Scale

0 50 100 km
800,000 800,000

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000

based on the areal distribution of the average where sufficient burial depths were reached. There,
geothermal gradient (Figure 5). All other Creta- the main source rock is the Gachetá (La Luna) For-
ceous and Tertiary strata are currently at tempera- mation, which reached maximum depths of up to
tures too low for hydrocarbon generation. Assum- 3000 m. The geothermal gradients in that area of the
ing current geothermal gradients and considering basin must have been in the range of 35°C/km and
the effective thickness of strata removed by higher for the source rocks to reach oil-window tem-
Holocene erosion near the fold and thrust belt, the peratures. Assuming the same matrix thermal con-
estimated temperature reached at maximum burial ductivity for rocks as exists at present, the effective
in the eroded part of the basin is insufficient to gen- thermal conductivity of the sedimentary succession
erate the hydrocarbons found currently in the basin at that time must have been at least 1.5 times higher
in Paleozoic, Cretaceous, Paleocene, and Eocene than today because of a significantly less proportion
strata. These hydrocarbons must have migrated of low-conductivity shales. This leads to paleoheat
into the present Llanos basin from sources located flow estimates for the early Tertiary of at least 80–90
to the west or have formed in the basin under pre- mW/m2. This thermal event coincides with the west-
viously higher temperature conditions caused by erly collision of a volcanic arc with the South
higher basement heat flow. American margin during Maastrichtian–Paleocene
The Ro levels are relatively low for their depths (McCourt et al., 1984), which led to back-arc high
and temperatures, particularly in post-Cretaceous heat flow similar to that found today in the Bolivian
strata, and suggest that the heat flow was also rela- Cordillera and Altiplano basin where an average heat
tively low during the late Tertiary. On the other flow of 84 mW/m2 has been observed (Henry and
hand, biodegraded oils are found in lower Tertiary Pollack, 1988). Heat flow has decreased since this
strata at depths greater than 3000 m. These oils must thermal event to reach the present relatively low val-
have been degraded in place prior to the middle-late ues. A possible reason for the rapid cooling could be
Miocene Andean orogeny, most probably during the that offered by Henry and Pollack (1988) for the sim-
late Oligocene–early Miocene, because only prior to ilarly low heat flow in the Oriente basin in Peru,
the early Miocene deposition of thick shaly strata southwest of the Llanos basin. A change in the sub-
(Leon Formation) was the Llanos basin open to duction of the Nazca plate and imbrication of cold
flushing by meteoric water. Only then would these oceanic lithosphere to the west may provide a mech-
oils have been at depths shallow enough to be at suf- anism for rapid reduction in heat flow during the
ficiently low temperatures for biodegradation to take late Tertiary to Holocene. This process, if real, may
place. For these oils to be biodegraded during the also explain the apparent eastward increase in heat
late Oligocene, they must have been generated from flow and geothermal gradients observed in the
Cretaceous rocks west of the present Llanos basin Llanos basin (Figure 5).
126 Geothermal Regime, Llanos Basin, Colombia

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000 Gauss


coordinates Figure 9—Estimated depth
Venezuela of the 120°C isotherm in
Colombia the Llanos basin (contour
N interval = 1000 m).
400

500
0
600 0

0
1,200,000 1,200,000

60 00
4000

00
500 0

30
40 00

3000

1,100,000 1,100,000
ra 0
d ille 600

or
C
n
er
st

00
Ea 5000

00
40
00

30
60

00
40

1,000,000 Bogota 1,000,000

00
50
00

900,000 00
30 900,000
40

30
00

Ba s in
5 00
0
of Guyana Shield
e
0
3000 g
Ed Scale
2 00

40
0 50 100 km
800,000 00
800,000

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000

Ro (%) A prior thermal event coincides with an Early


Cretaceous extension (failed rift) west to the pres-
0 0.2 0.4 0.6 0.8 1.0 1.2 ent Llanos basin in what is today the Cordillera Ori-
0 ental. The westerly position of the Cretaceous heat
source is also indicated by the westward R o
200 increase in Cretaceous strata (Figure 8). The pre-
Cretaceous thermal history of the basin is more dif-
400 ficult to evaluate because of a lack of information
regarding the burial history, thickness, and litholo-
gy of the eroded strata. A break at the top of the
600 Paleozoic in the Ro variation with depth suggests
that large packages of sedimentary rocks have been
800 eroded (e.g., estimated at 1500–2500 m in the well
of Figure 10). The high Ro values at the top of the
Depth ( m )

900 Paleozoic succession (Figure 11) show that maxi-


mum temperatures were reached by the Paleozoic
rocks prior to the deposition of the present post-
1000 Paleozoic succession. Otherwise, the old, pre-Cre-
taceous Ro levels would have been obliterated by
1200 subsequent higher temperatures, resulting in a con-
tinuous R o increase across the sub-Cretaceous
1400 unconformity. Unlike post-Paleozoic thermal events
Tertiary that took place to the west, the thermal event
Paleozoic responsible for Paleozoic source rock maturation
1600 took place to the southwest of the present basin, as
indicated by the Ro distribution at the top of the
1800 Paleozoic (Figure 11). This thermal event could be
related to a Triassic–Jurassic back-arc extension
2000 east of a magmatic arc, which formed as a result of
eastward-dipping subduction along the western
Figure 10—Vertical vitrinite reflectance profile in well margin of South America and North America (Pin-
Almagro-1, Llanos basin. dell et al., 1988). There is no known record of earlier
Bachu et al. 127

1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000 Gauss


coordinates Figure 11—Distribution of
Venezuela vitrinite reflectance in the
Colombia strata at the top of the
N Paleozoic succession,
Llanos basin (contour
1,200,000
1,200,000
interval = 0.2%). Dots rep-
resent data points.

1,100,000 1,100,000
ra
d ille
or
C
rn
ste dar
y
n
Ea Bou
zoic
leo
1,000,000 Bogota Pa 1,000,000

8
4

6
0

0.

2
1.

0.

1.
1.

4
0.

900,000 900,000

Ba s in
of Guyana Shield
e

g
Ed Scale

0 50 100 km
800,000 800,000
1,000,000 1,100,000 1,200,000 1,300,000 1,400,000 1,500,000 1,600,000

thermal events except, of course, for the early Paleo- event was probably caused by the collision of a vol-
zoic event corresponding to the failed rift that ini- canic arc with the western South American margin.
tiated the basin. The Early Cretaceous and early Ter- This event is likely responsible for the generation of
tiary thermal events most likely led to pulses in hydrocarbons in the Early Cretaceous graben west
hydrocarbon generation. Another episode of oil of the present Llanos basin. These hydrocarbons
generation is probably taking place today. As for migrated into the present Llanos basin prior to the
older episodes of oil generation, we cannot assess uplift of the Cordillera Oriental during the late
their occurrence because we lack additional infor- Miocene Andean orogeny. The heat flow in the
mation. basin then decreased from values estimated to be at
least 90 mW/m2 and higher to the moderate-to-low
values of today (average 42 mW/m2). One possible
CONCLUSIONS reason for this cooling could be imbrication of cold
oceanic lithosphere to the west associated with the
The general history of the Llanos basin and, in subduction of the Nazca plate.
particular, vitrinite reflectance distributions indi- The present geothermal regime in the Llanos
cate that during the basin’s evolution a series of basin is characterized by a generally westward-
thermal events took place that inf luenced the decreasing basement heat f low from values of
hydrocarbon generation in the basin. The basin 50–60 mW/m2 in the east near the Guyana Precam-
started in early Paleozoic as a failed rift. There is no brian shield to 30–40 mW/m2 in the west near the
information regarding the Permian–Carboniferous fold and thrust belt. Several local heat-flow highs in
period, which likely represents a hiatus in deposi- the south are associated with basement highs and
tion and perhaps considerable erosion. Little is probable granitic intrusions. The distribution of
known about the basin history until the Triassic– average geothermal gradients reflects the variation
Jurassic, when subduction along the western mar- of basement heat flow. Other factors also contribute
gin of South America produced a magmatic arc. A to the variation in average geothermal gradients
thermal event (high heat f low), most probably from 50°C/km in the east to 20°C/km in the west.
associated with this magmatic arc, is likely respon- One factor is related to increased thermal conduc-
sible for the relatively high maturation levels of tivity of the sedimentary column in the west
Paleozoic source rocks. Another failed rift event because of compaction in the deeper strata. Anoth-
with associated high heat flow took place west of er factor contributing to the observed geothermal
the present Llanos basin during the Early Creta- pattern could be cooling by westward downdip
ceous. A Late Cretaceous–early Tertiary thermal f low of formation water in Cretaceous–Eocene
128 Geothermal Regime, Llanos Basin, Colombia

aquifers in the central and western parts of the Henry, S. G., and H. N. Pollack, 1988, Terrestrial heat flow above
basin. We cannot ascertain the relative importance the Andean subduction zone in Bolivia and Peru: Journal of
Geophysical Research, v. 93, p. 153–162.
of these factors at this stage, but they should form Maze, W. B., 1984, Jurassic La Quinta Formation in the Sierra de
the objective of future quantitative studies of fluid Perijá, Venezuela-Colombia, and adjacent basins: geology and
and heat flow in the Llanos basin. tectonic environments of redbeds and volcanic rocks, in W. E.
The post-Paleozoic rocks are currently at maxi- Bonini, R. B. Hargraves, and R. Shagam, eds., The Caribbean–
South American plate boundary and regional tectonics: Geologi-
mum burial depth, except for a narrow area close cal Society of America Memoir 162, p. 263–282.
to the fold and thrust belt where they are slightly McCollough, C. N., 1987, Geology of the super giant Cano Limon
shallower. Because of the low Holocene heat flow, field and the Llanos basin, Colombia, in M. K. Horn, ed., Trans-
only the lowermost Cretaceous strata near the fold actions of the Fourth Circum-Pacific Energy and Mineral
and thrust belt are presently in the stage of oil gen- Resources Conference: Circum-Pacific Council, p. 299–316.
McCourt, W. J., T. Feininger, and M. Brook, 1984, New geological
eration, as indicated by vitrinite reflectance data and geochronological data from the Colombian Andes: conti-
and the depth of the 120°C isotherm. nental growth by multiple accretion: Journal of the Geological
Society of London, v. 141, p. 831–845.
Muñoz, F., 1991, El Paleozoico en la Cuenca de los Llanos Orien-
REFERENCES CITED tales. Futuro objectivo exploratoria, V: Exploración Petrolifera
en Cuencas Sub-Andinas, Tomo I.
AAPG, 1976, Basic data file from AAPG Geothermal Survey of Nyblade, A. A., and H. N. Pollack, 1993, A global analysis of heat
North America: University of Oklahoma, Norman. flow from Precambrian terrains: implications for the thermal
Bachu, S., 1985, Influence of lithology and fluid flow on the tem- structure of Archean and Proterozoic lithosphere: Journal of
perature distribution in a sedimentary basin: a case study from Geophysical Research, v. 98, p. 12207–12218.
the Cold Lake area, Alberta, Canada: Tectonophysics, v. 120, Peñalosa, M. P., and L. H. Ramirez, 1993, Determinación de histo-
p. 257–284. rias de compactación, paleopresiones de poro, y evolución ter-
Bachu, S., 1988, Analysis of heat transfer processes and geothermal mal—implicaciones en el modelamiento de cuencas sedimenta-
pattern in the Alberta basin, Canada: Journal of Geophysical rias: Universidad Industrial de Santander, Bucaramanga, 64 p.
Research, v. 93, p. 7767–7781. Pindell, J. L., S. C. Cande, W. C. Pitman III, D. B. Rowley, J. F.
Bachu, S., 1993, Basement heat flow in the Western Canada sedi- Dewey, J. Labrecque, and W. Haxby, 1988, A plate-kinematic
mentary basin: Tectonophysics, v. 222, p. 119–133. framework for models of Caribbean evolution: Tectonophysics,
Bachu, S., and R. A. Burwash, 1991, Regional-scale analysis of the v. 155, p. 121–138.
geothermal regime in the Western Canada sedimentary basin: Pollack, H. N., and D. S. Chapman, 1977, On the regional variation
Geothermics, v. 20, p. 387–407. of heat flow, geotherms, and lithospheric thickness: Tectono-
Bachu, S., and S. Cao, 1992, Present and past geothermal regimes physics, v. 38, p. 279–296.
and source rock maturation, Peace River arch area, Canada: Rybach, L., 1981, Geothermal systems, conductive heat flow,
AAPG Bulletin, v. 76, p. 1533–1549. geothermal anomalies, in L. Rybach and L. J. P. Muffler, eds.,
Bodell, J. M., and D. S. Chapman, 1982, Heat flow in the north-cen- Geothermal systems: principles and case histories: New York,
tral Colorado Plateau: Journal of Geophysical Research, v. 87, John Wiley, p. 3–76.
no. 84, p. 2869–2884. Rybach, L., 1988, Determination of heat production rate, in R.
Brigaud, F., D. S. Chapman, and S. Le Douaran, 1990, Estimating Haenel, L. Rybach, and L. Stegena, eds., Handbook of terrestrial
thermal conductivity in sedimentary basins using lithologic data heat-flow density determination: Dordrecht, Kluwer Academic
and geophysical well logs: AAPG Bulletin, v. 74, p. 1459–1477. Publishers, p. 125–142.
Bullard, E. C., 1947, The time necessary for a borehole to attain Tissot, B. P., and D. H. Welte, 1984, Petroleum formation and
temperature equilibrium: Monthly Notes of the Royal Astronom- occurrence, 2d ed.: New York, Springer-Verlag, 699 p.
ical Society, v. 5, p. 127–130. van der Kamp, G., and S. Bachu, 1989, Use of dimensional analysis
Chapman, D. S., T. H. Keho, M. S. Bauer, and M. D. Picard, 1984, in the study of thermal effects of various hydrogeological
Heat flow in the Uinta basin determined from bottom hole tem- regimes, in A. E. Beck, G. Garven, and L. Stegena, eds., Hydro-
perature (BHT) data: Geophysics, v. 49, p. 453–466. geological regimes and their subsurface thermal effects: Ameri-
Deming, D., 1989, Application of bottom-hole temperature correc- can Geophysical Union Geophysical Monograph 47, p. 23–28.
tions in geothermal studies: Geothermics, v. 18, p. 775–786. Villegas, M. E., S. Bachu, J. C. Ramon, and J. R. Underschultz, 1994,
Deming, D., and D. S. Chapman, 1988, Heat flow in the Utah- Flow of formation waters in the Llanos basin, Colombia: AAPG
Wyoming thrust belt from analysis of bottom-hole temperature Bulletin, v. 78, p. 1843–1862.
data measured in oil and gas wells: Journal of Geophysical Vitorello, I., V. M. Hamza, and H. N. Pollack, 1980, Terrestrial heat
Research, v. 93, no. 13, p. 657–672. flow in the Brazilian highlands: Journal of Geophysical
Deming, D., J. H. Sass, A. H. Lachenbruch, and R. F. De Rito, 1992, Research, v. 85, p. 3778–3788.
Heat flow and subsurface temperature as evidence for basin- Willet, S. D., and D. S. Chapman, 1986, On the use of thermal data
scale ground-water flow, north slope of Alaska: Geological Soci- to resolve and delineate hydrologic flow systems in sedimentary
ety of America Bulletin, v. 104, p. 528–542. basins: an example from the Uinta basin, Utah, in B. Hitchon,
Dengo, C. A., and M. C. Covey, 1993, Structure of the Eastern S. Bachu, and C. M. Sauveplane, eds., Hydrogeology of sedimen-
Cordillera of Colombia: implications for trap styles and regional tary basins: application to exploration and exploitation: Dublin,
tectonics: AAPG Bulletin, v. 77, p. 1315–1337. Ohio, National Water Well Association, p. 159–168.
Bachu et al. 129

ABOUT THE AUTHORS

Stefan Bachu Mauricio E. Villegas


Stefan Bachu received an engi- Mauricio Eduardo Villegas is a
neering degree in hydraulics, an staff research engineer at the Insti-
M.Sc. degree in hydrogeology, and tuto Colombiano del Petróleo (ICP)
a Ph.D. in transport processes from in Colombia. His activities are
the Technion-Israel Institute of directed to numerical simulation
Technology. After doing research of fluid flow in porous media. He
at Cornell University on a Lady holds a B.Sc. degree in civil engi-
Davis Fellowship, he joined the neering from the Universidad Jave-
Alberta Research Council (ARC) in riana in Colombia and an M.Sc.
Canada in 1983. In 1987, he be- degree in petroleum reservoir
came head of the petroleum geo- engineering from Texas A&M Uni-
science section, Alberta Geological Survey, ARC. His versity. After concluding the Llanos basin study, he
research interests include fluid flow, heat transfer and returned to Texas A&M University where he is current-
transport in sedimentary basins, modeling basin evolu- ly pursuing a Ph.D. in petroleum reservoir engineering
tion, and reservoir analysis. and simulation.

Juan C. Ramon Jim R. Underschultz


Juan Carlos Ramon is a geologist Jim Underschultz received his
in the exploration and exploitation B.Sc. degree in geology and M.Sc.
division of the Instituto Colom- degree in geodynamics from the
biano del Petróleo (ICP) in Colom- University of Alberta. He was with
bia. His main areas of interest are the Alberta Research Council from
geology, stratigraphy, organic geo- 1986 to 1994. Currently, he is a
chemistry, and source-rock poten- hydrogeologist with Hydro Pe-
tial in sedimentary basins. Current- troleum Canada, Ltd., in Calgary.
ly, he is pursuing an M.S. degree at His interests include hydrodynam-
the Colorado School of Mines. ics and mass transport in geological
environments, numerical simula-
tions, basin-scale tectonic processes, and electronic
mapping and visualization.

Vous aimerez peut-être aussi