Вы находитесь на странице: 1из 29

Tectonophysics 400 (2005) 179 207

www.elsevier.com/locate/tecto

Interpreting fracture development from diagenetic mineralogy and


thermoelastic contraction modeling
Renee J. PerezT, James R. Boles
University of California, Santa Barbara, Geological Sciences Department Room 1006, Webb Hall, Santa Barbara,
CA 93106-9630, USA
Received 1 March 2004; accepted 1 March 2005
Available online 9 April 2005

Abstract

Four sets of thin-section scale, Mode I (open mode), cemented microfractures are present in sandstone from the Eocene
Misoa Formation, Maracaibo basin, Venezuela. The first set of microfractures is intragranular (F1), formed early during
compaction and are filled with quartz cement precipitated at temperatures equal to or higher than 100 8C. The second set of
microfractures (F2) is cemented by bituminitepyrite, formed at temperatures between 60 and 100 8C, and are associated with
kerogen maturation and hydrocarbon migration from underlying overpressured source rocks. The third set of microfractures
(F3) is fully cemented by either quartz cement or calcite cement. The former has fluid inclusion homogenization temperatures
between 149 and 175 8C. These temperatures are mostly higher than maximum burial temperatures (~160 8C), suggesting that
upward flow, caused by a pressure gradient, transported silica vertically which crystallized into the fractures. Upward
decompression may have also caused a P CO2 drop, which, at constant temperature, allowed simultaneous carbonate precipitation
into the third microfracture set. The fourth set of thin-section scale microfractures (F4) is open or partially cemented by siderite
hematite and other iron oxides. The presence of hematite and iron oxides in microfractures is evidence for oxidizing conditions
that may be associated with the uplift of the Misoa formation. In order to time and place constraints on the depth of formation of
the fourth set of microfractures, we have coupled published quartz cementation kinetic algorithms with uniaxial strain equations
and determined if, in fact, they could be associated with the uplift of the formation. Our results suggest that thermoelastic
contraction, caused by the formations uplift, erosion, and consequent cooling is a feasible mechanism for the origin of the last
fracture set. Hence, we infer that meteoric water invasion into the fractures, at the end of the uplift, cause the precipitation of
oxides and the transformation of siderite to hematite.
D 2005 Elsevier B.V. All rights reserved.

Keywords: Sandstone diagenesis; Fracture formation; Veins; Thermoelastic contraction; Overpressure

T Corresponding author. Current affiliation: University of Calgary, Department of Geology and Geophysics, 2500 University Drive,
Northwest, Calgary, Alberta T2N 1N4 Canada.
E-mail addresses: rene@earth.geo.ucalgary.ca (R.J. Perez), boles@geol.ucsb.edu (J.R. Boles).

0040-1951/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2005.03.002
180 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

1. Introduction connection between sandstone horizons. The Misoa


Formation was deposited during Eocene time in a
Reservoir and thin-section scale partially cemented shallow marine-deltaic environment (Gonzalez et al.,
fractures and veins represent examples of tensile 1980), buried to more than 4 km under high
failure and fluid/fracture interactions. Extension sedimentation rates, subject to temperatures higher
fractures exhibit simple power-law scaling across than 180 8C (Rodriguez et al., 1997), and uplifted
3.44.9 orders of magnitude (Marret et al., 1999), but more than 2/3 of its maximum burial depth. Because
regardless of their scale, hydraulic fractures decrease of the wide range of pressure and burial regimes, the
pore fluid pressure in the adjacent rock. Furthermore, formation is ideal for the recognition of compaction,
depending on their connectivity, frequency, and and the study of inversion related fractures (Law and
hydrologic regime they may facilitate mass transfer Spencer, 1998). Previous diagenetic studies (e.g.,
and heat advection. Over time, some fractures are Ghosh et al., 1990; Perez et al., 1999b) and burial-
cemented while others remain open, causing a host- thermal reconstruction (Rodriguez et al., 1997; Perez
rock scale porosity and permeability heterogeneity et al., 1999a) of the Misoa Formation provide input
that ultimately affects fluid production. Microfrac- for our semi-quantitative thermo-mechanical analysis
tures are used to determine far field stresses (i.e., used to interpret fracture conditions during burial and
Laubach, 1989) and magnitude, direction, and scale uplift.
of fluid movement (e.g., Laubach, 1988; Eichhubl Our study describes the diagenesis of the Misoa
and Boles, 1997). Geochemical studies of cemented Formation sandstone and the mechanisms of fracture
fractures in sandstone are typically performed in development and cementation during burial and
tectonically deformed areas such as fault zones and uplift. The timing of cementation and fracture
folds (e.g., Hippler, 1993; Macaulay et al., 1997; forming-filling events are interpreted from fluid
Perez and Boles, 2004; Boles et al., 2004), but inclusion data and the application of the quartz
microfractures and veins are not restricted to precipitation kinetic model of Walderhaug (1996).
deformed areas. Based on petrography and diagenetic cements we
In the absence of tectonic stresses, compaction distinguish microfractures that originated during
disequilibrium and quartz cementation can result burial subsidence (fractures sets F1 to F3) from
fluid retention and pore pressure increase (Helset et others possibly originated during uplift (fracture set
al., 2002; Swarbrick et al., 2002), which may lead, in F4). The origin of the fourth fracture set is difficult
turn, to hydraulic fracturing (e.g., Laubach, 1988; to ascertain. Combining, however, petrologic infor-
Engelder and Lacazette, 1990). Uplift, erosion, and mation and diagenetic-mechanical modeling we
cooling can also cause significant pore fluid changes infer they have been formed and cemented in recent
and thermoelastic stresses that subsequently lead to in time at shallow depth, and are associated with
tensile failure (e.g., Suppe, 1985; Swarbrick and uplift.
Osborne, 1998). Tensile failure produces, in turn, Samples were collected from the Lagunillas,
pore pressure changes, allowing fluid movement, and Bachaquero, and Motatan onshore oil fields and a
leading, ultimately, to vein formation (Eichhubl and wildcat exploration well, BA-1, which is located
Behl, 1998). In these cases, the formation and approximately 2 km to the northwest of the Bach-
cementation history of veins indirectly reflect the aquero field (Fig. 1). No information on core or
host rocks stress history, fluid pressure, and fluid sample orientation was available. These localities
composition. provide, however, excellent chronology of fracture
The Misoa Formation, the subject of our study, is events. All fractures cements are monomineralic,
present in the subsurface on the east side of the which led us to group the structures in sets, based
Maracaibo Lake, Venezuela. The Misoa Formation is strictly on their mineralogy, depth of occurrence,
one of the most prolific siliciclastic hydrocarbon orientation relative to bedding, and cross-cutting
reservoirs of the world (Higgs, 1996), and, in this relations. The data and results that we herein present
paper we show microfractures provide avenues for may characterize the general diagenesis, veining
aqueous and hydrocarbon flow as well as vertical conditions, paleofluid composition, pore pressure,
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 181

Caribbean Sea

N
Venezuela

MARACAIBO

LAGUNILLAS
FIELD
BACHAQUERO

ER
FIELD

IV
R
A'

A
O
IS
M
MARACAIBO LAKE
A

MOTATAN
FIELD

Zulia Oriental Region

0 10 20 30 km

Fig. 1. Schematic map of the Maracaibo Lake basin depicting the Zulia Oriental Region and locations of oil fields mentioned in text. Well
coordinates and locations are proprietary information of Petroleos De Venezuela (PDVSA).

and stress regimes present during the burial and uplift ceous marine, and PaleogeneNeogene delta
of similar siliciclastic formations in other regions. sediments (Fig. 2; Lugo and Mann, 1995), derived
from cratonic sources as well as Caribbean and
Andean orogenic belts (Castillo et al., 1996). Late
2. Geologic setting Eocene and Miocene inversions, resulting from the
continental collision between the northern South
The studied sections of the Lower to Middle American border and the Lara Napes, interrupted the
Eocene Misoa Formation are located in the eastern subsidence of the basin (Lugo, 1991). During the
onshore region of the Lake Maracaibo basin, western initial stage of the collision (~54 Ma) the eastern
Venezuela (Fig. 1). The basin represents a stable onshore region, also called the Zulia Oriental Region
intracratonic terrain surrounded by active plate boun- (ZOR; Fig. 1), developed into a back-arc sub-basin,
daries and high relief areas (Gonzalez et al., 1980; that was quickly filled by turbidite, deltaic, and
James, 2000. The basin overlies a Mesozoic crystal- shallow marine deposits (Gonzalez et al., 1980). The
line basement and traps Jurassic volcaniclastic, Creta- Misoa Formation is a tidal-influenced deltaic deposit
182 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

Formations, Members, (Van Veen, 1972; Maguregui and Tyler, 1991; Higgs,
Age in Series and Zones
Million of 1996). Based on spontaneous potential and resistivity
MIO-PLEISTO-

Years Ago
logs, the sandstone to shale ratio is ~1:6. Several
PLIOCENE

(Ma). BETIJOQUE

ISNOTU normal and reverse faults cut through the Misoa strata
into Cretaceous strata, at more than 4 km subsurface
?
depth (Fig. 3). Reverse faults were generated by the
LA ROSA LAGUNILLA

Bachaquero
Lagunillas reactivation of CretaceousPaleocene normal faults
MIOCEN

Laguna (Roure et al., 1997).


La Rosa The burial history of the Misoa Formation consists
Santa Barbara of two subsidence periods, the first from 50 to 42 Ma
39.5-36
Regional unconformity and the second from 17 to 5 Ma, both interrupted by
EOCENE

to MENE GRANDE /
UPPER

39.
JARRILLAL uplifts. Much of the basin was inverted during the first
PAUJI Regional Seal uplift between ~42 and 38 Ma. In the study area, the
42.
EOCENE
MIDDLE

44
second uplift occurred from 7 to 4 Ma, and was
associated with the Miocene Orogeny. Vitrinite
49.
MISOA Studied Formation
reflectance data indicate a variable heat flow through-
EOCENE

out the history of the basin (Rodriguez et al., 1997).


EARLY

51. The thermal data are interpreted to represent a heat


flow increase from 52 to 58 mW/m2, resulting in
54 TRUJILLO thermal gradients between 30 and 35 8C/km from 54
to 49 Ma, followed by an exponential decay to ~48
CRETACEOUS PALEOCENE

66 GUASARE mW/m2, from 52 Ma to present. The present day


thermal gradient varies from 25 to 28 8C/km
(Gonzalez et al., 1980). The thermal interpretation
84 MITO JUAN/COLON Regional Source Rocks has been calibrated against apatite fission track,
LA LUNA
present-day temperature data, and temperature logs
(Rodriguez et al., 1997). Erosion plays an important
Fig. 2. Schematic stratigraphic column of the Zulia Oriental Region, role in our stress analysis and varies from 0.2 to ~3
western part of the Maracaibo basin. Source rocks underlies the km (Rodriguez et al., 1997). Based on measurements
Misoa Formation, whereas regional shale (seal) overlies (Gonzalez from exploration and production wells the Misoa
et al., 1980).

Well #
A LB-114 A'
S-SW N-NE
0 km

3.5 km
21 km

Undifferenciated
Trujillo Fm. Misoa Fm. Pauji Fm. La Rosa-Lagunillas-
Betijoque Fms.

Fig. 3. Schematic northeastsouthwest structural section through Tertiary strata of the studied area. The cross section location is depicted in Fig. 1.
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 183

Formation is currently at hydrostatic or sub-hydro- change original inclusion volumes, distorting tem-
static pressure due to extensive oil production perature interpretations (c.f., Goldstein and Rey-
(Gonzalez et al., 1980). nolds, 1994; Roedder, 1984). Methane inclusions are
found with H2O-rich inclusion assemblages. Thus,
we did not apply pressure corrections to the
3. Methods homogenization temperatures (Th) because small
CH4 concentrations cause Th to approximate true
3.1. Analytic trapping temperatures (Hanor, 1980). We calculated
salinity values from melting temperatures (Tm)
We collected approximately 70 sandstone core- assuming an H2ONaCl system and using Bodnars
plug samples along 300 m of core taken from seven (1992) revised equation of state. We used to Ths
wells, distributed over three oil fields and a wildcat (1) to delimit (in time) the precipitation of quartz
well in the east side of the Maracaibo Lake basin, overgrowth in intragranular fractures, quartz over-
western Venezuela (Fig. 1). The samples belong to growth surrounding detrital grains, quartz filling the
unfaulted segments of the Eocene Misoa Formation set of microfractures F3 using burial histories and
and their present depths range from 0.2 to 4.2 km, assuming the quartz filling occurred during burial,
which covers the entire formation (Fig. 2). From the and (2) to verify the quartz cementation modeling
70 core plugs we made 40 thin sections and stained results.
them for calcium carbonate with alizarine red and
counted 300 points per slide.
We observed and analyzed the fracture mineralogy 4. Modeling methods with brief theoretical
with a conventional polarized microscope and a JEOL background
scanning electron microscope (SEM) model JSM-
6300v. The SEM was set to an acceleration voltage of After determining the distribution and quantity of
15 kV and a beam current of 160 AA. An energy quartz overgrowths, we modeled in 1-D the precip-
dispersive analyzer (EDA Tracor Northern TN2010) itation of quartz overgrowth and the stress history of
was coupled with the SEM, and used a beam current the Misoa Formation. The quartz overgrowth model
of 250 AA. Additionally, the amount of quartz yields the amount of quartz cement and porosity
overgrowth was estimated in 29 quartz arenites and decline curves through time.
in 5 fractured sandstone samples with a cathodolumi- Our stress analysis yields the effective horizontal
nescence (CL) unit attached to the SEM. We obtained stress through time as a function of burial depth,
15 images per sample, covering an area of ~1 cm2, while changing the rock properties. The bridge
and determined quartz overgrowth percentages from between the diagenetic and stress model is the
digital analysis of CL images. porosity evolution. Experiments and empirical obser-
The elemental composition of calcite cements is vations suggest that mineralogy and porosity control
based on electron microprobe analysis using a the thermo-mechanical properties of sandstones (i.e.,
CAMECA SX50, set with an acceleration voltage of see review by Giles, 1997). In the Misoa sandstone,
15 kV, and beam current and beam diameter of 10 nA compaction and quartz cementation mainly control
and 10 Am, respectively. Results were normalized to 1 the porosity (Perez et al., 1999b). Thus, if the porosity
mol total of Ca + Fe + Mg + Mn. is calculated through time, depth, and temperature,
Homogenization and melting temperatures of the horizontal effective stress, assuming the uniaxial
fluid inclusions in quartz were determined using a stress model, is indirectly a function of time, depth,
Fluid-Inc., modified USGS gas-flow stage. Fluid and temperature.
inclusions were present in quartz overgrowth, quartz
grain-scale fracture cement, and in quartz veins in 4.1. Quartz cementation modeling
sandstone away from fault zones. We only used
fluid inclusions assemblages with constant liquid/ The quartz cementation model is based on numer-
vapor ratio to avoid necking-down effects, which ical algorithms, which consist of empirically derived
184 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

kinetics based on observations in the North Sea basin 2). When the T(t) path of the sandstone is known, T can
(Walderhaug, 1994), surface area (Walderhaug, 1996) be divided in small linear time steps. Thus, the total Vq
and a compaction decline function (c.f. Lander and precipitated from time t 0 to t n is the integration of Eq.
Walderhaug, 1999). This model has been tested in (1) along the Tt path, that is:
basins around the world, including the Maracaibo Z ti1
basin (Awwiller and Summa, 1997; Perez et al., M X n
Vq Ai 10btci1 di dt 3
1999a; Chatellier and Perez, 2000). The results of q i0 ti
these tests indicate that the use of the kinetic
parameters published by Walderhaug (1994) and where c is the temperature rate change or slope (8C/s)
presented in Table 1 yield errors smaller than 3%, and d i the initial temperature within the time step. The
which is acceptable to us. The algorithms were surface area A is estimated as the cumulative surface
numerically solved in a Microsoft Excelk spread- area of spheres of diameter D with the total volume
sheet. The basic concepts and mathematical formula- equal to the quartz fraction f times a unit volume of the
tions of the model are described elsewhere sandstone v (Walderhaug, 1996). The surface area A is
(Walderhaug, 1994, 1996; Walderhaug and Bjbrkum, readjusted after each time step and is considered
2003). Hence, we include here only a brief summary. proportional to the porosity loss through time and
The volume of quartz cement Vq (in cm3) precipi- depth, expressed as:
tated in a cm3 of sandstone with quartz surface area  
6f v /i
A (cm2), during time t (in s) at constant temperature Ai 1  coat 4
is: D /0

Vq M rAt=q 1 where / i is the porosity for the present time step and
/ 0 is the initial porosity at the time of deposition. The
where M is the molar mass of quartz (60.09 g/mol), r is porosity / i is also readjusted each time step because it
the rate of quartz overgrowth precipitation in mol/cm2 is considered a function of the effective vertical stress
s, and q is the density of quartz (2.65 g/cm3). The rate r through a compaction function (see Eqs. (5) and (6))
is expressed as a logarithmic function of temperature and the amount of quartz cement. The effect of grain
(Walderhaug, 1994): coating (coat parameter in Eq. (4)), which inhibits
cementation (Lander and Walderhaug, 1999), was
r a10bT 2
considered zero at all times.
where T is temperature (8C) and a and b are constants Different compaction decline functions have been
with units of mol/cm2 s and 1/8C, respectively (Table coupled with the quartz overgrowth precipitation

Table 1
Selected list of samples with thin-section scale microfractures and microveins
Well BA-1 Well LS-1387 Well LB-273
Depth (ft) Depth (m) Fracture type Depth (ft) Depth (m) Fracture type Depth (ft) Depth (m) Fracture type
1336 404.8 F4 1023.8 310.2 F2F5 6468.2 1960.1 F3
2503 758.5 F2F4 6359.4 1927.1 F2F5 6878 2084.2 F3
2506 759.4 F2 7340 2224.2 F5 7504 2273.9 F3
3327 1008.2 F2F4 9093.3 2755.5 F3F4 7729 2342.1 F3
3587.6 1087.2 F2 9343.3 2831.3 F3 7904.3 2395.2 F3F4
4169 1263.3 F2F4 9346.7 2832.3 F3F4 8890 2693.9 F2F3
4748 1438.8 F2F3F4 9842.8 2982.7 F2F3
5274 1598.2 F3F4
6345 1922.7 F2F3
Under cathodoluminescence, all samples contain F1 intragranular microfractures.
F2: bituminitepyrite microveins.
F3: quartz or calcite microveins.
F4: open uncemented microfractures or partially cemented by iron oxides.
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 185

Table 2
Parameters used in the quartz cementation kinetic model and thermo-mechanical parameters used in the stress analysis of sandstones
Symbol Variable Units Initial Final Source
value value
a Pre-exponential parameter, Eq. (2) mol/cm2 s 1.98e22 T Walderhaug (1994)
b Exponential parameter, Eqs. (2) and (3) 1/8C 2.2e2 T Walderhaug (1994)
C 12 Exponential constant, Eq. (5) 1/MPa 8.3e3 T Chuhan et al. (2002)
C 22 Exponential constant, Eq. (5) 1/MPa 8.5e3 T Chuhan et al. (2002)
C 32 Exponential constant, Eq. (5) 1/MPa 8.4e3 T Chuhan et al. (2002)
E4 Youngs modulus, Eq. (10) MPa4 1.0e3 16.5e3 Engelder (1985)
v4 Poissons ratio, Eqs. (7)(10) D 1.5e1 3.3e1 Bachrach et al. (2002),
Engelder (1985)
a4 Coefficient of thermal expansion, Eqs. (8) and (10) 8C1 10.0e6 10.8e6 Engelder (1985)
b Biot poroelastic parameter, Eq. (7) D 1 5.6e1 Breckels and van Eekelen (1982)
T Constant value through time.
1, 2, 3
Exponential constants for IGV reduction as a function of effective stress for fine, medium, and coarse-grained sandstones, see Eq. (5) in
text.
4
Assembled from various sources by Engelder (1985).
D, dimensionless.

model (e.g., Lander and Walderhaug, 1999). We used 4.2. Stress history modeling
new experimental compaction data based on effective
stress, rock composition, and grain size (c.f. Chuhan Cementation or mineralization change the thermo-
et al., 2002). The advantage of using the new data set mechanical properties of sediments (Jizba and Nur,
is two-fold. First, the data reveal a correlation between 1990; Giles, 1997). Experimental results demonstrate
intergranular volume decrease and grain size not that the Poissons ratio, Young modulus, poroelastic
revealed in other compaction studies (i.e., Pittman coefficient, and the coefficient of thermal expansion
and Larese, 1991). Second, the mineralogy of the vary significantly with mineralogy, porosity, and
sands used in these experiments is similar to the sandstone hardness (Engelder, 1985). Thus, because
mineralogy of the Misoa Formation; both are quartz sediments are unconsolidated when buried, diagenesis
arenites. We numerically fit Chuhans et al. (2002) ensures that the stress path during burial is different
results with curves of the form: from that during uplift.
Our model assumes that sandstone layers are
/ C1 eC2 rVv 5 poroelastic, homogenous, and isotropic. Thermo-
mechanical properties are assumed to (1) vary linearly
where C 1 and C 2 are constants (Table 2) and r vV is the and synchronously with the amount of quartz cement
effective vertical stress in MPa. The latter term, r vV, is and porosity throughout the burial; (2) attain their
the total overburden stress r v minus the pore fluid maximum values as porosity decreases to zero; and
pressure P f (Terzaghi and Peck, 1948). A cross plot of (3) remain constant after the minimum porosity is
predicted vs. measured amount of quartz cement for approached and during uplift (Warpinski, 1989).
this specific area using the kinetic parameters pre- Moreover, we assume that (4) the horizontal stress is
sented in Table 1 is shown elsewhere (in Fig. 3 from the minimum principal stress; (5) the overburden is
Perez et al., 1999b). the maximum principal stress; and (6) that all stresses
With the purpose of determining the origin of the are governed by the uniaxial strain model. Burial
fourth fracture set, which based on diagenetic cements histories, described before, are basic input for the
is inferred to have been created recently at shallow stress analysis.
depth, we use a simple thermoelastic model to In reality, our assumptions are not strictly met for
understand the possible stress evolution of the Misoa several reasons. A sandstone body may not behave
Formation. homogeneously, isotropically, and elastically due to
186 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

compaction, diagenesis heterogeneities, and stress to the point where microfractures occur. The temper-
concentration around flaws. Moreover, some micro- ature change as a function of depth Z may be
fractures may reflect regional tectonic stresses (Lau- expressed as:
bach, 1989), such as the late Eocene and Miocene
dT
plate convergence. However, based on the correlation DT Dz: 9
of fracture formations and model predictions, our dz
oversimplifications do not alter first-order effects of If the PTt path (burial history) of a sandstone
burial and uplift on the modeled state of stress (Narr body is divided in small linear time steps, the change
and Currie, 1982; Apotria et al., 1994). In the in the horizontal effective stress Dr hV can be calculated
Maracaibo basin, furthermore, the present day hori- numerically as a function of time. The sum can be
zontal stresses are low and vary significantly from one numerically performed along a PTt as:
oil field to another (Breckels and van Eekelen, 1982) Xn Xn 
suggesting that some oil field sub-regions, particularly vi
rhVi1 rvi  bi Pf i
those distant from major fault and folds, may behave i0 i0
1  vi
  
as tectonically relaxed areas. Ei
The mathematical formulations for the uniaxial V
ai Ti1  Ti rhi 10
1  vi
strain model are described elsewhere (e.g., Engelder
and Lacazette, 1990; Twiss and Moores, 1992). where i is a time step and (i + 1)  i is the linear
Therefore, we will only explain briefly the theory increment. As explained above, E, v, and a are
and numerical treatment. Assuming poroelastic assumed to vary linearly and synchronously with
behavior the effective vertical stress r vV in the uniaxial compaction, quartz cement, and ultimately the poros-
strain model is: ity, whereas b varies with Eq. (7). Breckels and van
Eekelen (1982) measured values of b on rocks of the
rvV rv  bPf 6
Misoa Formation. The initial and final values of E, v,
where b is the Biot poroelastic parameter (Biot, 1941) and a were taken directly from the literature and were
and P f is the pore fluid pressure. The model is based obtained experimentally from sandstone of similar
on the premise that the stress is partially accommo- composition, but different formations (see references
dated by elastic grain contacts. Empirical observations in Table 2). Additional stress terms in Eq. (10), for
suggest that exponential functions of the porosity example those that would account for tectonic
reproduce b (Giles, 1997), such as: stresses, are not included because they are not defined
 0:33 in the uniaxial strain model. The pore fluid pressure P f
/i varies with time. In the following section the
bi 7
/0 diagenetic evolution of quartz cement in the sandstone
where / 0 is the initial porosity at the time of of the Misoa Formation is calculated, including its
deposition and / i the porosity for the present time effect on porosity.
step. The horizontal stress (r h) is a function of the
overburden (Engelder and Lacazette, 1990). When the
effective horizontal stress includes the thermal effect, 5. Host rocks paragenesis and thermo-mechanic
r hV is calculated as: evolution
 
v E The Misoa sandstone is very fine, to medium-
rhV rv  bPf a DT 8 grained, well to moderately sorted quartz arenite to
1v 1v
sublitharenite. The average composition is quartz86-
where v is Poissons ratio, a is the coefficient of feldspar5lithic9 (Table 3). Early diagenesis in these
thermal expansion in 8C1, DT is the change in sandstones includes less than 1% of tangential grain
temperature, and E is Youngs modulus in MPa, the coating of smectite clay minerals, intergranular
horizontal effective stress r hV sometimes decreases framboidal pyrite, siderite cement, feldspar over-
faster than r vV during uplift, and may become tensile, growths, and less than 2% to 5% of kaolinite.
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 187

Table 3
Selected petrographic results
Well Depth Quartz Feldspar Met. Sed. Muscovite Kaolinite Calcite Siderite Pyrite S/I Chlorite QC Bituminite P. P S. P
(m) lithics lithics
BA-1 692.4 53 6 1 2 2 0 29 3 0 1 0 2 1 0 1
BA-1 878.3 59 4 5 1 2 0 7 1 0 13 0 4 0 0 3
BA-1 942.8 60 6 4 3 0 1 3 2 0 10 0 6 0 1 3
BA-1 1463.9 52 5 1 3 0 0 14 11 0 0 3 3 5 0 3
BA-1 1593.9 63 4 2 2 1 2 1 0 0 0 0 17 0 8 1
LB-114 2757.6 68 1 1 1 0 0 24 0 0 1 1 0 0 0 1
LB-1387 1392.4 60 2 2 1 1 0 28 0 1 0 0 4 0 0 0
LB-1387 2244.9 58 1 3 2 0 5 1 0 0 0 2 28 0 0 1
LB-1387 2412.4 60 2 11 3 1 1 2 1 0 0 2 15 0 0 3
LB-1387 2857.3 59 1 0 2 4 0 5 3 0 0 13 10 3 0 0
LB-1387 2858.3 57 2 3 1 1 0 2 6 0 0 12 15 0 0 1
LB-1387 3854.4 55 3 7 0 1 0 4 0 1 1 0 26 0 0 0
LB-273 1978.0 53 3 3 3 4 0 7 2 0 3 6 12 1 1 0
LB-273 2040.1 67 4 1 1 0 1 0 1 0 1 0 12 1 10 0
LB-273 2100.6 56 5 6 1 4 1 0 0 2 1 5 16 0 0 2
LB-273 2103.4 59 4 6 1 1 0 3 2 0 0 3 17 3 0 0
LB-273 2293.1 61 2 6 2 2 0 11 3 0 0 0 8 1 0 3
LB-273 2363.6 49 11 6 2 3 0 1 2 0 5 0 16 4 0 1
LB-273 2511.9 56 4 8 2 2 1 3 2 1 6 15 1 0 0 0
LB-273 2694.8 61 1 5 1 2 1 0 1 0 0 0 27 0 0 0
LB-273 2718.7 62 6 4 1 4 0 2 3 0 0 8 6 1 0 3
LS-837 496.1 61 8 2 4 1 0 21 0 0 0 0 3 0 0 0
The relative % is based on 300 point counts per thin section.
Met. lithicsmetamorphic lithic fragments.
Sed. lithicssedimentary lithic fragments.
S/Ismectite/illite clay.
QCquartz cement as overgrowth.
P. Pprimary porosity.
S. Psecondary porosity.

Petrographic observations suggest that kaolinite is tion temperatures from two-phase liquid/vapor iso-
related to feldspar dissolution and predates quartz lated fluid inclusions, present within the last
cement. overgrowth event, ranged from 100 to 175 8C,
Intermediate diagenesis is primarily characterized averaging 119 8C (Table 4, Fig. 4). Melting temper-
by quartz overgrowth formation. Quartz overgrowths atures vary from 0.4 to 2.2 8C, yielding salinity
postdate the aforementioned minerals and are volu- values between 0.8 and 2.7 wt.% NaCl, assuming an
metrically the most important authigenic phase in the H2ONaCl system (Bodnar, 1992). Modeling results,
host rock. Overgrowths average 12% of the bulk rock, based on Walderhaug (1996) algorithms, indicate that
but in fine-grained sandstone may reach 26% of the the precipitation of quartz overgrowth took place at
bulk host rock (Table 3), based on cathodolumines- depths between 2.5 and 4 km. This range (2.5 to 4
cence (CL). Petrographic and CL observations indi- km) corresponds to 10 and 18 my after deposition,
cate the presence of two distinct generations of quartz mostly during early Eocene to Oligocene time (i.e.,
cement. The first quartz cement generation is volu- Fig. 5). Based on fluid inclusion data solely it would
metrically less than 2% and may be inherited (Ghosh be tempting to presume that quartz cement also
et al., 1985). The second generation of quartz cement precipitated during uplift, however, according to the
has a darker luminescence than the first, contains fluid modeling results, quartz cement does not precipitate in
inclusion assemblages away from dust rims, and is the the sandstone beyond 18 my after deposition, even
main factor reducing the rock porosity. Homogeniza- though it is within the precipitation window, because
188 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

Table 4
Typology of fluid inclusion data analyzed on sandstones
Well Field Depth P.D.T M.B.T. Occurrence Th range Th average Tm range Tm wt.%
(m) average NaCl
LB-1182 Bachaquero 1180.3 64 117 Grain-scale quartz 111 to 117 (n = 11) 113.5 F 2 1 to 1.6 2.74
cemented 2.3 (n = 2)
microfracture (F1)
MOT-4 Motatan 2545 63 N 63 Grain-scale quartz 98119 (n = 11) 105.6 F 6.5 N.D. N.D. N.D.
cemented
microfracture (F1)
LS-1257 Lagunillas 3819.1 131 170 Quartz microvein (F3) 149175 (n = 14) 167.6 F 6.5 0.4 to 0.9 1.57
1 (n = 13)
LS-1257 Lagunillas 3819.1 131 170 Quartz overgrowth 145160 (n = 4) 151.3 F 6.3 0.4 to 0.7 1.23
1 (n = 2)
MOT-4 Motatan 2545 63 N 63 Quartz overgrowth 107129 (n = 6) 116.3 F 7.5 1.6 to 0.5 0.88
2.2 (n = 5)
BA-1 Wildcat 1953 48 N 48 Quartz overgrowth 101128 (n = 5) 110.2 F 10.4 N.D. N.D. N.D.
MOT-2 Motatan 2565 68 N.D. Quartz overgrowth 106139 (n = 11) 113.7 F 9.3 1 to 1.3 2.24
1.8 (n = 5)
Temperatures in Celsius.
P.D.T.present day temperature in Celsius.
M.B.T.maximum burial temperature based on thermal modeling.
Thhomogenization temperature.
Tmmelting temperature.
wt.% NaClassuming watersalt system, and using Bodnar (1992) revised equation.
N.D.not determined.

there is no intergranular space available for quartz quartz cement model, are clean, quartz-rich, and with
cement to grow. no other cement than quartz overgrowths.
Locally, post-quartz siderite and calcite cement The diagenetic model suggests that the porosity in
filled up to 18% of the intergranular volume, and the clean quartz arenites decreased to almost zero
prevented further compaction and quartz cementation. before the maximum burial depth was attained (Fig.
Based on microprobe analysis, the average composi- 5C,D). Experimentally, it has been demonstrated that
tion of the late-stage calcite is Ca0.91Mg0.01(Fe + the Poissons ratio v, Youngs modulus E, and
Mn)0.08CO3. The precipitation PTt of the local poroelastic parameter b of sandstone in quartz-rich
late-stage calcite may be inferred using results from sands vary with the porosity and degree of cementa-
the quartz precipitation kinetic model (Fig. 5), specif- tion (see review by Giles, 1997). The mathematical
ically, using intergranular volume (IGV) reduction function describing how these properties vary with
curves through geologic time. The method consists of porosity and time, i.e., either exponentially or linearly,
matching the available IGV left by compaction and is to our knowledge unknown. Consequently, we
quartz overgrowth with the amount of late-stage calcite arbitrarily made v, E, and a of clean quartz arenites to
cement quantified by microscopic analysis. Post-quartz evolve linearly and synchronously with the porosity
calcite cement is calculated to have precipitated and amount of quartz cement, from their initial values
between 100 and 150 8C, at depths greater than 3 km, at deposition to their final values when / c 0 (Fig.
between 12 and 14 my after deposition. These petro- 5A,B). For instance, the porosity of clean quartz
graphic results and cementation temperatures are arenites in well LB-273 at 2488 m depth decreasedas
similar to other studies in the basin (e.g., Chatellier a function of quartz cementation and compaction
and Perez, 2000). Other local late-stage cements from 47% to less than 2% from 0 my to 18 my after
represent 2% to 6% of the host rock and consist of deposition (Fig. 5C,D), respectively. Consequently,
siderite, chlorite, and chert (Table 3). However, the the v, E, and b varied from 0.15 to 0.33, from
majority of the sandstones, in which we apply the 1  103 to 16.3  103, and from 1 to 0.56,
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 189

HOMOGENIZATION TEMPERATURES
(in fractures and host rocks)
A B
7 7
n= 21 LS-1257, 3819 m n= 26 MOT-4, 2545 m
6 Veins (F3) 6 Microfractures (F1)
Overgrowth Overgrowth
5 5
Frequency

Frequency
4 4

3 3

2 2

1 1

0 0
<140 145 150 155 160 165 170 175 180 180 90 95 100 105 110 115 120 125 130 135 140 145 150
Th (oC) Th (oC)

ICE MELTING TEMPERATURES


(in fractures and host rocks)
C D
8 8
n= 17 LS-1257, 3819 m n= 5 MOT-4, 2545 m
7 Veins (F3) 7 Microfractures (F1)
Overgrowth Overgrowth
6 6
Frequency

5 5
Frequency

4 4

3 3

2 2

1 1

0 0
-1 -0.9 0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 -2.4,-2.2 -2.0-1.8 -1.6 -1.4 -1.2 -1.0-0.8 -0.6 -0.4 -0.2 0.0
Tm (oC) Tm (oC)

Fig. 4. A and B show fluid inclusion data from quartz-filled microfractures (set F1 and F3) from the Misoa Formation. Quartz precipitation
occurred over a wide temperature range, beginning at 100 8C. (C, D) Ice melting temperatures suggesting the presence of marine-diluted and
meteoric water influence in Misoa pore waters throughout the burial history.

respectively, during the same period of time (Table 2; 6. Analysis of microfracture morphology
Fig. 6AD). The maximum burial depth was reached
20 my after deposition (Fig. 5A), 2 my after all rock Four kinds of thin-section scale fractures occur in
properties had attained their maximum values. Misoa sandstone. The lack of core orientation forced
In our model the rock properties vary with time, us to group the fractures in sets strictly based on their
temperature, and diagenesis, not with depth solely, mineralogy, cross cutting relations, and orientation
and after the sandstone is totally cemented, when relative to lamination. Intragranular fractures repre-
porosity reaches zero, the mechanical properties sent the earliest set (F1) and are filled with quartz
remain constant. Understanding the evolution of the cement. The second set of thin-section scale micro-
aforementioned parameters is important because they fractures (F2) are filled with bituminitepyrite and are
are part of the basic input for our stress analysis in truncated by a third set of microfractures (F3) filled by
determining the host-rocks thermo-mechanical con- either quartz or calcite cement. The fourth set of
ditions during uplift. microfractures (F4) is either uncemented or partially
190 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

BURIAL - THERMAL HISTORY OF HOST ROCKS


0 0
500 A Present day
depth 20 B Present day
temperature
1000 40

Temperature (C)
1500 60
80
2000
Depth (m)

100
2500
120
3000
140
3500
160
4000 180
4500 200
0.0 10.0 20.0 30.0 40.0 50.0 0.0 10.0 20.0 30.0 40.0 50.0
Time after deposition (MY) Time after deposition (MY)

MODEL RESULTS: QUARTZ CEMENTATION AND COMPACTION OF HOST ROCK


30 50
C 45 D Thin section
porosity based
25 Measured on point count
amount of 40
quartz 35
Primary porosity %
Quartz cement %

20 overgrowth
30
15 25
20
10
15
10
5
5
0 0
0.0 10.0 20.0 30.0 40.0 50.0 0.0 10.0 20.0 30.0 40.0 50.0
Time after deposition (MY) Time after deposition (MY)

Fig. 5. A and B represent burial and thermal history for fine-grained, clean, quartz-rich, and quartz cemented sandstone from well LB-273, at
2488 m depth (Rodriguez et al., 1997). (C) Quartz cementation kinetic model results, based on Walderhaug (1996) algorithms (Eqs. (1)(5)),
representing the 26% of quartz cement precipitated through time. (D) Calculated porosity decreases through time as a result of compaction and
quartz cement.

cemented by hematite and other concurrent iron Milliken and Laubach, 2000; Chuhan et al., 2002). All
oxides. The latter cement type and set of micro- intragranular fractures are Mode I, formed by a
fractures (F4) are observed in sandstone strata uplifted displacement normal to the fracture wall. However,
more than 1/3 from their maximum burial depth. All they vary randomly in pattern, shape, and size. These
microfractures are Mode 1 (open mode), extension fractures generally have straight or curvilinear traces
fractures, and confined to host rocks rich in inter- (Fig. 7), they cut grain boundaries, and in some cases
granular quartz cement. Their morphology and aper- cut overgrowth cements. Displaced fractured walls
tures are described in the following text. indicate normal dilation, followed by solid-volume
increase, and repacking. In some cases, microfractures
6.1. Intragranular fractures (F1) extend across two or three grains forming trans-
granular microfractures. Wedge-shaped fractures at
A summary of the characteristics of the micro- graingrain contacts commonly display triangular
fractures is presented in Table 5. Intragranular shapes in a radiating pattern, and are diagnostic of
fracturesfractures within or across grainsare the stress concentrations (Laubach, 1997). The radiating
most common type of fracture in the Misoa Formation patterns are most likely produced by compaction with
and represent post-depositional features associated or without tectonic stresses present (Milliken and
with compaction during burial (e.g., Milliken, 1994; Laubach, 2000). Many quartz-filled intragranular
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 191

0.35 0
A -2000
B
0.3
-4000

Young's Modulus E
Poisson's Ratio V 0.25
-6000

(MPa)
0.2 -8000

0.15 -10000

-12000
0.1
-14000
0.05
-16000

0 -18000
6.0 11.0 16.0 21.0 26.0 31.0 36.0 41.0 46.0 51.0 0.0 10.0 20.0 30.0 40.0 50.0 60.0
Time after deposition (MY) Time after deposition (MY)

0.000012 1.2
Coefficient of Thermal Expansion

C D
0.00001 1

Biot Poroelastic
0.000008 0.8

Parameter
(1/C)

0.000006 0.6

0.000004 0.4

0.000002 0.2

0 0
0.0 10.0 20.0 30.0 40.0 50.0 0.0 10.0 20.0 30.0 40.0 50.0
Time after deposition (MY) Time after deposition (MY)

Fig. 6. Synchronous and linear variation of v, E, and a with / through geologic time (see Eqs. (1)(5)); b varies with Eq. (7). Example for
quartz arenites from well LB-273, presently at 2488 m depth; dashed line represents the approximate time of maximum burial depth. We assume
the parameters evolve from their initial values at the time of deposition, to the final values when the primary porosity reached zero / c 0 (Table
1), which is before maximum burial depth. Results are basic input for thermoelastic contraction calculations. The final b value corresponds
specifically to sandstones from the Misoa Fm. The rest are theoretical values taken from the literature, see Table 2.

fractures are inherited, but these, however, have sets based on their mineralogy and relative timing of
distinctive bright luminescence under CL and blunt formation, not orientation as lack of oriented cores
terminations at grain boundaries, which distinguishes prevented us from determining fracture direction and
them from quartz-filled fractures developed in situ from further testing our grouping method. Never-
in the present setting. Based on image analysis of 15 theless, the three sets, in order of occurrence, consist
different CL photographs from three samples (five of (1) microfractures partially or fully cemented by
photographs per sample) the quartz cement trapped in bituminitepyrite (F2), (2) microfractures fully
microfractures averages less than 3% of the total cement by either quartz or calcite (F3), and (3)
photograph. Qualitatively, there is a positive correla- microfractures either open or cemented by hematite
tion between the abundance of intragranular micro- iron oxides (F4). All fracture filling cements display
fractures and depth, similar to that observed in the finer crystals than the host rock, except for fractures
Gulf Coast basin (Makowitz and Milliken, 2003). filled by calcite cement. The following three sets of
microfractures can be described as separate events,
6.2. Cemented thin-section scale fractures (F2F4) because their mineralogy suggests that they formed at
different geochemical conditions and geological time.
Other microstructures present in Misoa sandstone
are microfractures that extend across thin sections. 6.2.1. Second set of microfractures (F2)
They occur either isolated or in parallel sets. Again, The microfractures of the second set are filled by
we grouped these microstructures in three different bituminite (residual oil) and pyrite cement (Fig. 8A
192
Table 5
Characteristics of vein generations including timing, fracture mechanisms, and possible temperature range
Basin event Fracture Microvein Microvein Fracture Orientation Host rock Mechanism Fluid flow, and Temperature

R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207


generation cements secondary cement termination relative intergranular that lead to mass transport (8C)
to bedding cement fracturing mechanism
lamination
Subsidence F1 Quartz Quartz cement Triangular, Random Quartz arenites Compaction Grain scale silica ~ 100+
tabular, Stress diffusion
radiating
pattern
Oil generation F2 Bituminite Pyrite, siderite Triangular Sub-parallel Quartz cemented CD, kerogen Kilometer scale from 50100TT
migration-charge quartz arenites maturation underlying source rocks
Maximum burial F3 Quartz None Abrupt/blunt High angle Quartz cemented CD, kerogen Local diffusion or less 140170+
quartz arenites maturation than a kilometer scale
upward flow
Maximum burial Calcite None Abrupt/blunt High angle Quartz cemented CD, kerogen Local diffusion or less 140170+
quartz arenites maturation han a kilometer scale
upward flow
Uplift F4 Uncemented Pyrite, Hematite Triangular Random, Quartz cemented Thermal Small scale diffusion b 50T
high angle quartz arenites stresses and meteoric water influx
Temperatures derive from fluid inclusion data.
T Hydrocarbon generationmigration from Rodriguez et al. (1997).
+
Inferred temperature from IGV of host rock, petrology, and burial-thermal curves.
TT Inferred from petrology and thermoelastic contraction modeling.
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 193

QG 6.2.2. Third set of microfractures (F3)


The third set microfractures is filled by quartz and
calcite cement, they are orthogonal or cut at high
QC angle to the set filled by bituminitepyrite (Fig.
8C,D). Quartz and calcite microfractures are present
in samples that are currently at the same depth range.
They are grouped in sub-parallel sets, at a high angle
F1
with respect to bedding lamination. They also post-
date quartz cement or occur within the intergranular
QG QC quartz cementation window. The microfractures of the
QG third set are 0.3 to 0.5 mm wide (aperture) and up to 2
to 3 cm long. In several samples, F3 microfractures
overlap, adjoining different segments. Fracture termi-
200 microns nations are abrupt, and the fracture spacing is
Fig. 7. Sample from well LB-1182, 1172 m depth; F1 intragranular relatively close, ranging from 1 to 3 per 2 cm2 of
quartz-filled microfractures; in black is quartz cement QC, in gray thin section.
the detrital quartz grain QG. The figure depicts the solid volume In core samples, calcite crystals in these micro-
increase, normal dilation, and stress concentration at grain contacts. fractures are euhedral. Under CL, the crystals are dull-
The triangular fracture shape suggests compaction origin but we
orange to non-luminescent and lack zonation. The
have no information on orientation.
sandstone adjacent to the fracture-filling space com-
monly consists of irregular zones of calcite cement,
0.1 to 0.2 mm in width. These rims may have formed
D). These fractures are 0.2 to 0.5 mm wide (aperture) by host rock displacement through filling by calcite,
and less than 0.5 cm long. They occur at a high indicating that part of the fracture cement grows
angle with respect to bedding lamination, in parallel outward into the host rock matrix, as well as inward
sets. They are sub-laminar and follow grain bounda- into the vein center. The euhedral crystal terminations
ries (Fig. 8B). These observations suggest that are indicative of growth into open space, whereas the
fracture propagation and bituminite infiltration were cements diffuse contact with the host rock is
both controlled by fluid percolation through a indicative of dissolution/precipitation or matrix
compacted, but not fully cemented, host rock. recrystallization.
Fracture terminations are triangular tips, depicting Quartz-cemented microfractures are sub-perpen-
simple tapering. Fracture spacing is usually less than dicular to the set F2, which is filled with bituminite
0.5 cm. Bituminite-filled microfractures have shrink- pyrite (Fig. 8D,E). Euhedral quartz crystal termina-
age cracks filled by quartz cement (Fig. 8B) tions indicate growth into open fracture space. Quartz
indicating quartz precipitation in the presence of cemented fractures occur along grain boundaries.
hydrocarbons. In the absence of quartz crystals, Under CL, the quartz cement in veins has dark
bituminitic microfractures are also filled by concur- luminescence and lacks zonation (Fig. 8F), indicating
rent siderite and pyrite. a diagenetic origin. We suggest that the mechanism
Based on petrographic observations, siderite cor- responsible for the F3 microfracture event took place
roded adjacent quartz grains and percolated in the after hydrocarbon migration, based on the fact that the
host rock adjacent to the vein. Based on SEM-EDX F3 quartz and calcite set cut F2 bituminitic micro-
pyrite, in contrast, precipitated within the bituminitic fractures.
filled space. The bituminite cement in microfractures
is in all cases less than 2% of the total rock. The 6.2.3. Fourth set of microfractures (F4)
presence of bituminitic microfractures suggests that The microfractures from the fourth set consist of
the fracture event took place after hydrocarbon transgranular, uncemented, and partially cemented
generation-expulsion or synchronously with hydro- microfractures (Fig. 9AC), that are isolated or cut
carbon charge. the previously described microfracture sets F2 and F3,
194 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

A D

Bituminite
microvein
F2

Bituminite filling Quartz-filled


microfractures microfracture
F2 F3
0.1mm 2mm

B E
Bituminite-filled
microfracture
F2 Quartz
bridge

Quartz-filled
microfracture
F3

0.02mm 2mm

C F
Bituminite-filled
microfracture
F2 Quartz-filled
microfracture
F3
Quartz overgrowth
Calcite-filled preceeding calcite
cement
microfracture
F3
Bituminite-filled
microfracture
F2

1mm 1mm

Fig. 8. Representative photomicrographs of microfracture sets F2F4; (A) sample from well LS-1387, 2755.4 m depth with microfracture filled
by bituminitepyrite (set F2); (B) detail of quartz filled cracks, same sample as above. (C) Sample from well BA-1, 1450 m depth, calcite
cemented microfracture (F3) postdating (cross cutting) bituminite filled microfractures (F2) intergranular quartz cement. D to F are from well
LB-1387, 2755.5 m depth, (D) F2 displaced by set F3; (E) conjugated microfractures filled by quartz cement (F3) in quartz arenite sandstone;
(F) cathodoluminescence reveals no zonation, possibly (but not conclusively) suggesting a single precipitation event in F3 quartz-filled
fractures.

and occur in sandstone that are presently buried to 0.5 cm long. Fracture terminations are blunt, triangu-
depths less than 0.5 km deep. These sandstone bodies lar tips, and cut grains, suggesting a high degree of
have been uplifted more than 2.5 km. These micro- cementation and cohesion of the host rock at the time
fractures are Mode I in origin, as the rest of the sets, of fracturing. Irregular crystals of siderite, pyrite, and
and are 0.2 to 0.5 mm wide (aperture) and less than patches of hematite line the fracture walls (Fig. 9B).
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 195

Furthermore, this set (F4) can also be associated with


the dilatancy and reopening of microfaults (Fig. 9C).
A Microfaults, when present in uplifted intervals of the
Misoa Formation, reveal dilatancy and reopening, and
have small sub-euhedral quartz crystals growing
inward into the opened space, suggesting a contrac-
tion event and late-stage precipitation.

Open uncemented
microfractures
7. Interpretation of fracture cements from
F4
petrologic, microthermometric, and probe data

7.1. Generation of the first set of microfractures

Fluid inclusions occur in quartz-filled, intragranu-


2mm lar microfractures. The inclusions occur in small
B
clusters without preferred orientation and contain
Iron Oxide filled
microfracture
two visible fluid phases at room temperature. In most
F4 cases, the homogenization temperature (Th) in intra-
granular microfractures is lower than the Th in
overgrowths. These observations clearly suggest that
P quartz grain fracturing and re-sealing preceded quartz
overgrowth (Fig. 10). Moreover, fluid inclusion data
suggest that quartz cement precipitated initially within
microfractures and subsequently surrounded grain
surfaces, as observed under CL in other studies
P (e.g., Milliken, 1994; Laubach, 1997; Makowitz and
Milliken, 2003). The observations also indicate that
2mm
fluid salinity remained relatively constant during
C quartz precipitation (Fig. 10).

Reopened 7.2. Generation of the second set of microfractures


Microfault
A compositional analysis of residual hydrocar-
bons present in microfractures is beyond the scope
of our research. Based, however, on burial-thermal
histories and the most current oil generation-migra-
P P
tion model, the Misoa Formation was charged with
hydrocarbon at temperatures between 60 and 100
8C (Rodriguez et al., 1997). Thus, bituminitepyrite
microveins represent a fracture event that occurred
within the same temperature range or later than the
0.5mm
oil generation-migration event. Furthermore, geo-
Fig. 9. Three representative F4 photomicrographs; (A) Well BA-1, chemical experiments demonstrate that when resins
1922 m depth uncemented tensile fracture; (B) well BA-1, 1008.2 m and asphaltenes are reduced between 60 and 100
depth, partially cemented fractures by iron oxides and pyrite; (C)
well LS-1387, 1925 m depth, microfaults are common features and
8C, they yield H2S, CO2, and viscous black tar (see
we interpret reopening associated to uplift, based on calculations in review by Machel, 1987), similar to that observed
text and cementation. P is pore space. in pipelines.
196 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

Ice Melting Temperature Tm (C)


-0.5

-1

nce
urre
-1.5 occ
of
pth
De
Legend
-2 Quartz cement in fracture set F1
Quartz overgrowth in grains
Quartz-filling microfractrures set F3
-2.5
90 100 110 120 130 140 150 160 170 180
Homogenization Temperature Th (C)

Fig. 10. Homogenization vs. ice melting temperatures from quartz cements in overgrowths, intragranular fractures, and quartz-filled
microfractures indicating the relative depth and sequence of occurrence.

7.3. Generation of the third set of microfractures and gases, which decreases P CO2. The P CO2 drop at
constant temperature increases the carbonate satura-
The third set of microfractures is filled by either tion, which can lead to carbonate precipitation. If the
quartz and or calcite. Calcite and quartz micro- Fe/Ca ratio of the pore fluid is relatively low (Boles
fractures are grouped together because they both cut and Ramseyer, 1987; Eichhubl and Boles, 1998)
bituminite-filled fractures in an orthogonal direction calcite would precipitate upon decompression, if high,
(Fig. 8C,D), but they do not occur in the sample nor at siderite would dominate. Petrographic study indicates
the same depth. Homogenization temperatures in that F3 calcite microveins formed after oil emplace-
quartz cements present in the third set of micro- ment, cutting F2 bituminitepyrite microveins (Fig.
fractures F3 range from 149 to 175 8C suggesting a 8C). Thus calcite and quartz filled microfractures may
late stage of fracturing, and quartz precipitation (Table
5). Matching these temperatures with burial histories,
it is possible that the cementation occurred after (Fe+Mn)CO3
5%
hydrocarbon emplacement, consistent with F3 cross
cutting of F2. Melting temperatures range between 1
and 0.4 8C (Fig. 10), yielding salinity values
between 1.7 and 0.5 wt.% NaCl assuming an H2O
NaCl system (Bodnar, 1992). The salinity range
indicates dilute marine and meteoric fluid sources
were present during F3 fracture cementation.
Based on microprobe analysis, calcite crystals in n=18
calcite microfractures averages Ca 98(Fe + Mn) 1.6
Mg0.4CO3, have a Fe/Mg ratio of about 3, and have
from 1 to 3 mol% Fe + Mg + Mn substitution for Ca
(Fig. 11). Deep late-stage carbonate precipitation has
been reported in the Misoa Formation (e.g., Perez et CaCO3 MgCO3
al., 1997) as well as in other basins of the world (i.e., 100% 5%

Boles 1998). In general, calcite becomes less soluble Fig. 11. Normalized molar compositions, based on microprobe
with increasing temperaturei.e., at greater depths. In analysis, of calcite cements present in microfractures showing
addition, fracturing causes decompression of fluids relatively pure composition.
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 197

have occurred within the same temperature range, i.e., took place after fracture formation (Fig. 10), at depths
~150 to 175 8C, even though they do not coexist in ~3 km corresponding to r vV from 25 to 30 MPa
the same sample. It is important to denote that some, assuming a ~25 8C/km thermal gradient and hydro-
but not all, fractures filled with calcite cement show static pressure gradient, respectively.
quartz lineaments or quartz grains growing into the Based on the shape, mode, fracture shape, and
fracture, predating the calcite cement (see detail in relative orientation to bedding (Table 5; Figs. 8 and 9),
Fig. 8C). This type of relationship could be associated we interpret microfractures F2 to F4 to be extensional
to synkinematic quartz cement, and possibly, post- and to have formed perpendicular to the minimum
kinematic calcite cement (Laubach, 2003). confining stress r 3. Based on their inferred depth of
formation and the absence of nearby faults, micro-
7.4. Generation of the fourth set of microfractures fractures F2, F3, and F4 may have formed when P f
was higher than r 3, so that the minimum effective
The fourth set of microfractures are transgranular, stress r 3V became tensile and higher than the fracture
extensional, Mode I, and are partially cemented by toughness To. In other words, generations F2 to F4
pyritehematite and other concurrent iron oxides. may have occurred hydraulically, meaning r 3V =
Pyrite is the main fracture cement, whereas hematite r 3b d P f b To, where To is the tensile strength of
is secondary but still fills the fracture space. The the rock (Fig. 13). The presence of P f causes a rock to
presence of late-stage hematite and iron oxides behave as if the confining pressure were decreased by
supports a diagenetic event associated with uplift an amount equal to P f (Twiss and Moores, 1992). At
and exposure to meteoric fluids. Hematite is stable in small differential stress, an increase in pore pressure
reducing environments with pH N 4.6, however, its drives the effective stress toward the tension field
free energy is greater under oxidizing conditions and along the normal stress axis, where it meets the
may result from oxidation of pyrite or siderite (Garrels criteria for tension failure (Fig. 13A). At large
and Christ, 1965). Pyrite and siderite have small differential stresses (large r 1Vr 3V), an increase of pore
stability fields, and should precipitate under strongly pressure would shift the stress toward the failure
reducing conditions at pH N 7 (Garrels and Christ, envelope in the Coulomb failure criteria, causing
1965). Thus, dissolution of pyrite or siderite and shear.
precipitation of hematite is a reaction that can take Hydraulic fractures by themselves are, by no
place under oxidizing conditions during influx of means, indicative of overpressures. However, within
meteoric water associated with uplifting of the Misoa our geologic context, burial depths greater than 3 km,
Formation at temperatures less than 50 8C. the presence of an oil-gas source rock, the hydraulic
microfractures F3 strongly, but not conclusively,
suggest the presence of paleo-overpressures, meaning
8. Microfracture and mass transfer mechanisms (a) the pore fluid ( P f) pressures was higher than
hydrostatic, and (b) the r 3V was higher than the
Compaction and repeated fracturing in the Misoa hydrostatic stress, higher than the fracture toughness,
Formation led to four microfracture sets (F1 to F4; and lower than the overburden stress.
Fig. 12; Table 5). Intragranular microfractures (set F1) Overpressure is created by a pore fluid volume
were the first set to occur and compaction stresses are increase with minimal change in porosity and at a rate
usually interpreted as the main fracture mechanism for that does not permit dissipation (Swarbrick et al.,
these (e.g., Milliken, 1994). Compaction experiments, 2002). Several mechanisms may have caused P f to
performed under hydrostatic pressures and using increase above hydrostatic in the Misoa Formation.
sands of similar composition and grain size to The first to occur in the burial sequence and most
Misoas, demonstrate that intragranular fractures important is compaction disequilibrium, which is
could initiate at 6 MPa and increase in number and caused by compaction exceeding rates of pore fluid
size at stress levels up to 2030 MPa (Chuhan et al., loss. Clay dehydration, smectiteillite transformation,
2002). Average Ths in F1 fractures are 105.6 F 6.5 kerogen transformation, gas generation, hydrocarbon
and 113 F 2 8C (Table 4) suggesting that cementation buoyancy, quartz cementation, and osmosis are other
198 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

Present
Time from deposition (MY)
day
0.0 10.0 20.0 30.0 40.0 50.0
0.0
A
500.0

1000.0

Depth (m) 1500.0

2000.0

2500.0

3000.0

3500.0

4000.0
4500.0

(1) Intragranular
microractures F1

(2) Quartz cement overgrowth


within intragranular fractures

(3)Quartz cement overgrowth


around detrital grains
FRACTURE PARAGENESIS

(4)Occurrence of set F2

(5)Bituminite filling fractures

(6)Occurrence of set F3

(7)Calcite cement ?
filling microfractures set F3

(8)Synkinematic quartz
overgrowth into fractures

(9)Postkinematic calcite
cement filling fractures

(10)Quartz filling completely


F3 microfractures
(11)Occurrence of set F4

(12)Meteoric water
invasion
(13)Iron Oxides fills set F4

Fig. 12. Paragenetic sequence of fracture events present in the Misoa Formation. *The occurrence of F4 contraction fractures varies from field to
field, depending upon the burial history. **Iron oxides precipitated in sandstones units that were uplifted to depths less than 0.5 km from the
surface. (1) Based on Chuhan et al. (2002) experiments; (2) based on fluid inclusions data (FI); (3) based on FI data and Walderhaug, 1996
model; (4) inferred from petrography; (5) based on timing of oil migration (Rodriguez et al., 1997); (67) inferred from petrography; (89) based
on quartz lineaments; (10) based on FI data and petrography; (11) based on thermoelastic contraction model; (1213) based on the presence of
iron oxides.

causes that occur at a relatively same temperature At depths shallower than 2 km there is generally
window (see review by Swarbrick and Osborne, little potential for overpressure generation. However,
1998). Several of the mechanisms above may have in basins characterized by sedimentation rates
occurred in isolation, simultaneously, or sequentially, between 0.5 and 1 km/my and restricted pathways
triggering overpressure and ultimately microfracture for fluid loss, such as the Central North Sea basin and
in the Misoa Formation. Nile Delta (Law and Spencer, 1998) as well as the
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 199

A

Shear
stress

Tension Compression
Mohr
rion
crite Envelope
c ture
m b fra
lo
Cou

Hydraulic fracture Effective stress Applied stress

Overburden
Tensile fracture
criterion
n 3= 1 3 1 3 1 ^n
Normal stress

Final state
of stress after Pf Initial state
increase Pf of stress

Pf increase leads to tensile failure

B
Shear
stress

Tension Compression
n
r iterio Mohr
ctu re c
b fra Envelope
Co ulom

Hydraulic fracture

Overburden
Tensile fracture
^

criterion
n 3= 1 3 1 3 1 3 1 ^n
Normal
stress
^

Initial state
of stress
Final state
of stress after cooling

Fig. 13. Mohr circle sketch diagram representing various possible stress stages of the Misoa Formation; (A) burial cycle. Pore fluid pressure P f
reduces the consolidation. Tensile fractures will form if the minimum stress is equal of higher than the fracture toughness, i.e., r 3 z To. During
burial, increasing pore pressure shifts the differential stress to the tensile field leading to hydraulic fracturing; (B) uplift cycle. During uplift the
initial state of stress, as a result of erosion and cooling, may increase the differential stress and shift it to the tensile field. As we have shown the
net effect of load decreased and thermal contraction could be extensional.
200 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

Maracaibo basin, pressure profiles reveal overpressure the pore pressure. At constant temperature, conse-
at depths shallower than 1.3 km. quently, the P CO2 drop after hydraulic fracturing
Clay dehydration also occurs in the Maracaibo increases carbonate precipitation. Moreover, if the
basin (Perez et al., 1997) and its effect may add to the temperature is decreased during upward flow, the
overpressure caused by compaction disequilibrium P CO2 decrease caused by vertical fluid drainage could
(Swarbrick and Osborne, 1998). As explained, outweigh the temperature effect, and also favors
increased pore fluid pressures shift the minimum carbonate precipitation into fractures (Boles and
effective stress r 3V to the tensile field at values equal Ramseyer, 1987; Eichhubl and Boles, 1998). Further-
to, or higher than, the fracture toughness To (Fig. 13A). more, hydraulic fracturing could occur under adiabatic
Microfractures filled by bituminitepyrite cement, (constant heat) irreversible conditionsdue to the low
formed at depths greater than 2 km, based on oil thermal conductivity of rocksleading to temperature
migration and generation models. We interpret these increases caused by the internal work generation of
microstructures to reflect hydraulic microfractures the fluid phase during decompression (Wood and
associated with overpressure caused by kerogen Spera, 1984). The increase in temperature, accompa-
maturation, oil expulsion/gas generation, and gas to nied by P CO2 drop, may decrease carbonate solubility
oil crackingin addition to disequilibrium compac- even further, leading to faster carbonate precipitation.
tion. Kerogen maturation was most likely absent in Additionally, geochemical experiments suggest
the Misoa Formation due to its low TOC, high O2 that shale in the Misoa and La Luna Formations
index, and low oil generation potential (Jaffe and release cyclic and acyclic organic acids at %Ro 0.33
Gardinali, 1990). Thermobarometry and organic 0.77 and %Ro 0.51.8, respectively (Jaffe and
maturation modeling, however, suggest the presence Gardinali, 1990). In general, the thermal breakdown
of near-lithostatic pressures in the underlying source of kerogen yields directly CO2, CH4, H2S, and organic
rock La Luna Formation (Fig. 2) during oil gener- acids (Lundegard and Land, 1986). At high concen-
ation/expulsion (Sweeney et al., 1995; Vrolijk et al., trations (i.e., N 1.4 M acetic acid at 60 8C) organic
1996a and 1996b). These high pore pressure con- acids act as a buffer to added CO2 resulting in
ditions may have been transferred upward to lower carbonate precipitation (Surdam and Crossey, 1985).
sections of the highly cemented Misoa. Upward At lower organic acid concentrations, carbonate would
transfer of pressure is a common mechanism and is be precipitated with P CO2 drops, presumably caused
a main control of overpressure distribution in basins during veining and its associated decompression and
such as the central North Sea and Mahakam Delta upward fluid flow (Eichhubl and Boles, 1998).
(Swarbrick and Osborne, 1996). Based on correlations between the amount of
Upward transfer of excess pore fluid pressure may quartz cement and concurrent fractures in rocks
have created a pressure gradient that could explain lacking any other tectonic driver (Laubach, 1988), it
hydrofracture, upward flow, and the addition of has been proposed that the rapid rate of porosity
dissolved carbonate and silica into microfractures. reduction due to quartz overgrowth precipitation may
Microfractures filled with quartz cement (F3) have be another primary factor in generating overpressures
Ths z maximum burial temperatures and contain (Helset et al., 2002). Given the high cementation rates
abundant oil fluid inclusions. Thus, they might in Misoas sandstones, it is not unreasonable to
indicate upward transport of hydrocarbons and silica, assume quartz cementation was an important factor
by decompressing fluids into fractures, from intervals in elevating fluid pressures. However, a much more
that were more advanced in burial diagenesis and investigation is needed on this particular topic. For
temperature. The quartz lineaments in some calcite- example, Swarbrick and Osborne (1998) disregarded
filled microfractures (Fig. 8C) strongly suggest the the importance of quartz cementation as an over-
presence of synkinematic quartz overgrowth and post- pressure generator based on Darcys Law.
kinematic calcite precipitation. The origin of open uncemented microfractures,
Calcite filling microfractures (F3) could also reflect such as the F4 set, is difficult to ascertain. Open
decompression and upward fluid flow. Fracturing microfractures may result from stress during drilling,
increases the total pore volume locally, decreasing core handling, sample preparation, or unloading
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 201

during coring (Nelson, 1981; Laubach, 1997). We initial values of v, E, and a (Table 2) are assumed
interpret some F4 fractures, however, as the result of constant during burial. Subsequently, after the max-
basin unroofing, similar to those described by imum burial depth is reached v, E, and a are at final
Walderhaug (1992) based on the presence of hematite values shown in Table 2, and remain constant there-
cement, which suggest a late stage of oxidation event. after. Results indicate that the horizontal stress is
Precipitation of hematite and iron oxides takes place compressive at maximum burial (r h 40 MPa) and
under oxidizing conditions, possibly during meteoric tensile during uplift (11 MPa). The fracture tough-
fluid invasion associated with uplift of the Misoa ness is generally between 5 and 20 MPa for
Formation. Present fluids of non-treated reservoirs in sandstone (Suppe, 1985), and is exceeded after ~1/2
the Misoa Formation have an Na/Cl N 1, TDS ~1500 of the thickness of the strata is gone (Fig 14A). Hence,
mg/l, and Ca/Mg ratio averaging 1.5, clearly suggest- the appropriate conditions for fracturing formed
ing meteoric influence (Vasquez, 1998). toward the end of the uplift. Subsequently, fractures
Moreover, a stress analysis of the formation during were filled with iron oxides at shallow subsurface
uplift supports our argument. The net effect of uplift, conditions, less than 0.5 km depth, and presumably at
which implies erosion and cooling, is that the hydrostatic fluid pressure conditions.
horizontal stress becomes the minimum compressive In our second approximation, equations for effec-
stress and the effective horizontal stress becomes tive horizontal stress (Eqs. (3)(10) are integrated
tensile even at depths greater than 1 km. We believe numerically throughout the timedepth path of the
that this mechanism, also called thermoelastic con- Misoa sandstone (Fig. 5A,B). Even though F2 and F3
traction, may have been an important factor during the microveins strongly suggest an excess of fluid
generation of the last microfracture event. pressures throughout the Misoa Formation during
deep burial, we assume minimum conditions for
fracturing, i.e., near-hydrostatic fluid pressures, where
9. Modeling results in support of thermoelastic k = P f /r v varied from 0.45 to 0.51. Clearly this
contraction assumption is an underestimation. Any increase in
P f over hydrostatic shall accentuate the stress effect,
To support our argument that late stage fractures leading to a faster or deeper thermoelastic contraction.
(F4) are associated with uplift, we analyzed the stress We assumed a Poissons ratio of 0.15 for uncon-
conditions for fracturing during uplift assuming a solidated sand (Bachrach et al., 2002) and 0.33 for
uniaxial strain model. In the uniaxial strain model, if well-cemented sandstone (Engelder, 1985). The bulk
the rocks behave elastically, the Poisson effect overburden density increases from 2 (rock + pores) at
predicts that a decrease in the vertical load causes the time of deposition to 2.65 g/cm3 (highly quartz
expansion in the vertical direction, and contraction in cemented sandstone) at maximum burial depth. Hence
the horizontal direction. Thus, the horizontal stress at the total vertical stress is greater during upliftdue to
any given depth may be quite different during burial higher rock densitythan at the corresponding depth
relative to uplift. during burial. The poroelastic parameter decreases
In our first simple approximation we calculated the from 1 to 0.56, following Eq. (6). As we explained, v,
effect of burial and uplift on stress on a representative E, and a are assumed to vary linearly and synchro-
PT curve of a Misoa horizon (Fig. 5A,B; Eq. (9)). nously with the porosity and degree of cementation
From the assumption of tectonic stress being absent, during burial (Fig. 6AC). Because the porosity
we assume the maximum and minimum horizontal approaches zero before the maximum burial depth,
normal stresses are equal (Dr H(max) = Dr H(min)), a the parameters remain constant at their final values
function of the vertical stress, and temperature. For before maximum burial depth is achieved and during
simplicity, we chose a geobaric and thermal gradient uplift (Fig. 6AD), although high fracture densities
of 25 MPa/km and 25 8C/km, respectively. The may change the bulk density. Various rock properties
calculations were done at three points: at the surface and burial history scenarios from wells MB-1, LB-
(0 m), at maximum burial (4 km), and at minimum 1387, LS-1257, LB-114, mostly around the Bach-
burial after uplift (1.5 km). In this calculation, the aquero field (Fig. 1) were incorporated into the stress
202 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

A Stress History Analysis


Effective Horizontal Stress (MPa)

-30 -20 -10 0 10 20 30 40 50 60 70 80 90 100


0
Tension Compression
Field Field
500
Calculation in 3 time steps

1000
Burial Depth (m)

1500
Buria

2000
l

2500

3000
Up

3500
lift

4000

B Effective Horizontal Stress (MPa)


-30 -20 -10 0 10 20 30 40 50 60 70 80 90 100
0
Tension Compression
Field Field
500
Calculation in 30 time steps. Based
on Equations 1 to 10.
1000 (See text)
Burial Depth (m)

1500
Buria

2000
l

2500

3000
Up
l
ift

3500

4000

Fig. 14. Effective horizontal stress as a function of burial depth assuming uniaxial strain model. (A) Calculation performed in three steps
assuming thermo-mechanical parameters constant through time (Table 2), based on Eqs. (8)(10). (B) Calculation performed in 30 time steps
assuming that thermo-mechanical parameters vary through time (Table 2), based on Eqs. (1)(10). Results from both calculations show that the
effective horizontal stress r hV becomes lower than P f at ~3.7 km depth and becomes tensile 3 Ma, at less than 2.5 km depth. Shaded area
represents typical values of fracture toughness To.
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 203

history calculation and consistently demonstrated that 10. Discussion


uplift reduces r hV to the tensile field (Fig. 13B).
It is noteworthy to recall that the Biot coefficient Models as simple as the one we are presenting
was originally defined to accommodate the elastic cannot express accurately the complex stress, burial or
deformation of mineral grains under pressure. How- pressure history of Misoa sandstone for various
ever, at temperatures above 100 8C the rocks volume reasons. For instance, observations from geologic
reduction is also partially controlled, among many structures and in situ measurements suggest that a
other parameters, by chemical compaction and not by state of stress in which r1 = r3 is uncommon (e.g.,
mechanical squeezing solely (Bbjrkum, 1996). Under Engelder and Geiser, 1980). Pressure-solution and
these circumstances the use of the Biot coefficient in fracturing change the basic assumption of elasticity
Eqs. (6)(10) perhaps would not adequately describe (Narr and Currie, 1982). Several faults that are present
the decreased effect of the fluid pressure on the in the studied oil fields can accommodate radial
effective vertical stress. Nonetheless, during burial, contraction of the strata during uplift, however there is
chemical compaction, although inelastic and irrever- no correlation between the occurrence of micro-
sible, will also decrease the net effect of the vertical fractures and these faults. And last, but not least, we
stress (rock overburden load) decreasing consequently neglected tectonic stress because its identification and
the effective vertical stress. In view of the above quantification is difficult, subjective, and prone to
discussion, we decided to perform simulations in error (Narr and Currie, 1982).
which the Biot coefficient was considered constant Thermoelastic contraction modeling has been used
and equal to 1 at all times. In other words, neglecting to explain fracture genesis in several sedimentary
its effect. The results are illustrated with dashed lines basins. For instance, Narr and Currie (1982) devel-
in Fig. 14B. All calculations seem to indicate that, if oped a stress-history model to explain joints in Utahs
the Biot coefficient is dropped or not considered, it Uinta basin. Similar to us, their elastic model included
would be easier to obtain a tensile horizontal effective effects of overburden, pore pressure, temperature, and
stress at the end of the uplift period, supporting further tectonic strains, and they suggested that high pore
our interpretation. fluid pressure was responsible for the joint system
Summarizing, under our assumptions, for increas- during uplift. Engelder (1985) studied loading paths
ing values of r hV during burial, P f exceeds r hV leading and joint propagation of the Appalachian Plateau
to conditions favorable for hydraulic fractures only if during different tectonic cycles including uplift. The
P f is above hydrostatic. However, during uplift, even study also concluded that failure propagated by
under hydrostatic pressure and at depths greater or contraction during uplift. Warpinski (1989) developed
equal to 1 km the effective horizontal stress is a mathematically complex model for estimating stress
ultimately reduced to the tensile field. The Poisson states in reservoirs for the Piceance basin assuming
ratio contributes to lower stresses, but the slope of r hV elastic and viscoelastic rock behavior. The model
during uplift is dominated by the Youngs modulus in included pore pressure, temperature gradients, and
the thermal stress term. In other words, thermal consolidation, but lacking, however, diagenetic
contraction of the rock associated with uplift over- effects. The model concluded that stresses in sand-
whelms the Poisson effect and leads to tensile failure. stone are fairly accurately represented with an elastic
In addition to the horizontal effective stress, the fluid analysis similar to ours. Apotria et al. (1994) studied
pressure must exceed the fracture toughness To in the fracturing and stress history of the Antrim Shale in
order to propagate a fracture (Fig. 14B). For the the Michigan basin, and concluded that unloading
purpose of this discussion, this term is not included in provided an important mechanism for fractures con-
our stress modeling. It is noteworthy that sandstone sistent with the contemporary stress field of the area.
fracture toughness is typically between 5 and 20 We believe that the calculations, herein presented,
MPa (Suppe, 1985), and r hV in our model reaches are useful in deciphering physical effects and geologic
between 5 and 25 at the end of the uplift, processes that are important in controlling the state of
depending upon the burial history analyzed, and the stress in quartz-rich sandstone during different stages
inclusion or exclusion of the Biot coefficient. of burial and uplift. We also believe that, in spite of
204 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

the oversimplifications, the model reproduces first suggesting that calcite may have a post-kinematic
order effects of burial and uplift. Finally, although we origin. As discussed previously, hydraulic F3 fractures
cannot prove that open uncemented microfractures by themselves are not indicative of overpressures. As
were caused by tensile stresses during uplift, they are we explained, however, within the Misoas geologic
consistent with our first order predictions of stress context, they suggest the presence of pressures above
conditions resulting from thermoelastic contraction hydrostatic.
within the Misoa Formation. We interpreted open uncemented and pyrite
hematite partially cemented microfractures (F4) to
have formed late in the diagenetic history as a result of
11. Conclusions meteoric water incursion during the late stage of
uplift. Based on our calculations, we conclude that the
Intragranular quartz-filled microfractures (set F1) Misoa Formation had conditions favorable for tensile
in the quartz-rich Misoa Formation are the most failure during the uplift. We also conclude that a total
common microstructures and represent post-deposi- denudation, uplift, or surface exposure does not
tional structures associated with compaction stresses. constitute a requirement to drive the stress to the
Ths in F1 fractures are lower than in overgrowths, tensile field. Contraction fractures can occur even at
indicating that quartz grain fracturing and annealing several kilometers depth, if tectonic stresses are
predates quartz overgrowth. Quartz overgrowth is absent. Thermal elastic contraction of sandstones
the most important authigenic phase in the fracture may have been the dominant, but not the only,
host rocks, averaging 12%, but may locally reach up mechanism responsible for late stage, fourth set of
to 26%. Quartz precipitated between 100 and 175 8C fracturing. Our analysis indicates that there is an
during burial. Modeling and fluid inclusion data intimate interplay between host-rock diagenesis, over-
suggest that the porosity in quartz arenite decreased pressure, hydraulic fracturing, mass transport, and
to almost zero before the maximum burial depth. veining mechanisms that should be examined as
Sequentially, four sets of microfractures formed in coupled transformation processes.
the Misoa Formation. Bituminitepyrite (set F2) as
well as quartz and calcite filled microfractures (set
F3) suggest periods of hydraulic extension that Acknowledgments
could be associated with overpressure. Compaction
disequilibrium in the Misoa may have been the We thank E. Perez and Dr. Arthur Sylvester for
primary overpressure mechanism, however, it may their editorial and critical reviews. Early comments by
have been augmented by an upward pressure Drs. Jean-Yves Chatellier, Santosh K. Ghosh, and
gradient generated by underlying oil source rocks Thomas L. Dunn were very useful. Suggestions by the
at the time of kerogen maturation and by the rate of journal referees Drs. Kevin Furlong, Olav Walder-
porosity reduction due to quartz cementation. haug, and Steven Laubach greatly improved the
Homogenization temperatures in quartz-filling micro- content of the paper. PDVSA-Exploration provided
fractures (F3) range between 149 and 175 8C, samples and the U.S. Department of Energy (DOE)
whereas maximum burial temperatures are around funded our research under Grant No. 444033-22433.
160 8C. The vertical upward transfer of pore fluid
pressure explains the formation and addition of silica
into F3 fractures from intervals more advanced in References
diagenesis. Microfractures filled by calcite (set F3),
which postdate those filled by bituminitepyrite, may Apotria, T., Kaiser, C.J., Caine, B.A., 1994. Fracturing and stress
reflect hydrofracture, decompression, and a subse- history of the Devonian Antrim Shale, Michigan basin. In:
quent P CO2 drop, which may have outweighed any Nelson, J., Laubach, S. (Eds.), Rock Mechanics. Balkema,
Rotterdam, pp. 809 816.
temperature effect on solubility, leading to carbonate Awwiller, D.N., Summa, L.L., 1997. Quartz cement volume
precipitation. Synkinematic quartz cement growing constraint on burial history analysis: an example from the
into the fracture space predates calcite cement, Eocene of Western Venezuela. American Association of
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 205

Petroleum Geologists and SEPM Annual Meeting, Abstracts, Appalachian plateau. Journal of Geophysical Research 85,
vol. 6, p. 66. 6319 6341.
Bachrach, R., Dvorkin, J., Amos, M.N., 2002. Seismic velocities Engelder, T., Lacazette, A., 1990. Natural hydraulic fracturing. In:
and Poissons ratio of shallow unconsolidated sands. Geo- Barton, N., Stephansson, O. (Eds.), Rock Joints. Balkema Press,
physics 65, 559 564. pp. 35 44.
Biot, M.A., 1941. General theory of three-dimensional consolida- Garrels, R.M., Christ, C.L., 1965. Solutions, Minerals, and
tion. Journal of Applied Physics 12, 155 164. Equilibria. Jones and Barlett Publishers, Boston. 450 pp.
Bodnar, R.J., 1992. Revised equation and table for freezing point Ghosh, S., Maguregui, J., Garcia, L., Aguado, B., 1985. Diagenesis
depressions of H2Osalt fluid inclusions (Abstract). PACROFI and quality of sandstones from the Misoa Formation, lower
IV. Fourth Biennial Pan-American Conference on Research on Eocene, north of the Maracaibo Basin. Congreso Geologico
Fluid Inclusions, Program and and Abstracts, pp. 108 111. Venezolano 1, 749 774.
Bbjrkum, P.A., 1996. How important is pressure causing dissolution Ghosh, S., Aguado, B., Maguregui, J., Di Croce, J., Isea, A., 1990.
of quartz in sandstones? Journal of Sedimentary Research 66 Meteroic diagenesis and porosity enhancement in fluvio-deltaic
(1), 147 154. reservoir facies; B-6 sands, Maracaibo Basin, Venezuela.
Boles, J.R., Ramseyer, K., 1987. Diagenetic carbonate in American Association of Petroleum Geologists and SEPM
Miocene sandstone reservoir, San Joaquin Basin, California. Annual Meeting Abstracts, vol. 74, p. 661.
American Association of Petroleum Geologists Bulletin 71, Giles, M., 1997. Diagenesis: A Quantitative Perspective: Implica-
1475 1487. tions for Basin Modelling and Rock Property Prediction. Kluwer
Boles, J.R., Eichhubl, P., Garven, G., Chen, J., 2004. Evolution of a Academic Publishers, Boston, MA. 526 pp.
hydrocarbon migration pathway along basin bounding faults: Goldstein, R.H., Reynolds, J.T., 1994. Systematics of Fluid Inclusions
evidence from fault cement. American Association of Petroleum in Diagenetic Minerals. SEPM Short Course, vol. 31. SEPM
Geologists Bulletin 88, 947 970. (Society for Sedimetary Geology), Tulsa, Oklahoma. 160 pp.
Breckels, I.M., van Eekelen, H.A.M., 1982. Relationship between Gonzalez, D.J., Iturralde, J., Picard, X., 1980. Geologia de
horizontal stress and depth in sedimentary basins. Journal of Venezuela y sus Cuencas Patroliferas. Ediciones Foninves,
Petroleum Technology, 2191 2199. Caracas. 1031 pp.
Castillo, M.V., Lugo, J., Ostos, M., 1996. Paleogene clast Hanor, J.S., 1980. Dissolved methane in sedimentary brines:
provenance in the eastern Maracaibo basin in the context of potential effect on the PVT properties of the fluid inclusions.
the Caribbean and Andean tectonics. American Association of Economic Geology 75, 603 609.
Petroleum Geologists Bulletin 80, 1279. Helset, H., Lander, R., Matthews, J., Reemst, P., Bonnell, L., Frette,
Chatellier, J.Y., Perez, R., 2000. Timing of fluid flow and mass I., 2002. The role of diagenesis in the formation of fluid
transfer revealed by numerical modeling of quartz and overpressures in clastic rocks. In: Koestler, A., Husdale, R.
calcite cementation: examples from Maracaibo Basin (abs). (Eds.), Hydrocarbon Seal Quantification. Special Publication-
American Association of Petroleum Geologists Annual Meet- Norwegian Petroleum Society, NPF, vol. 11, pp. 37 50.
ing, p. A26. Higgs, R., 1996. A new facies model for the Misoa Formation
Chuhan, F.A., Kjedstad, A., Bjbrlykke, K., Hbeg, K., 2002. Porosity (Eocene), Venezuelas main oil reservoir. Journal of Petroleum
loss in sand by grain crushingexperimental evidence and Geology 19, 249 269.
relevance to reservoir quality. Marine and Petroleum Geology Hippler, S.J., 1993. Deformation microstructures and diagenesis in
19, 39 53. sandstone adjacent to an extensional fault; implications for the
Eichhubl, P., Behl, R., 1998. Diagenesis, deformation, and fluid flow and entrapment of hydrocarbons. American Association of
flow in the Miocene Monterey Formation. In: Eichhubl, P., Petroleum Geologists Bulletin 77, 625 637.
Behl, R. (Eds.), Diagenesis, Deformation, and Fluid Flow in the Jaffe, R., Gardinali, P.R., 1990. Generation and maturation of
Miocene Monterey Formation. The Pacific Section SEPM, Long carboxylic acids in ancient sediments from the Maracaibo basin,
Beach, no. 83, pp. 5 13. Venezuela. Organic Geochemistry 16, 211 218.
Eichhubl, P., Boles, J., 1997. Scale and dynamics of fracture-related James, K.H., 2000. The Venezuelan hydrocarbon habitat: Part 1.
fluid flow in the Miocene Monterrey Formation, coastal Tectonics, structure, palaeography and source rocks. Journal of
California. Geofluids II 97, pp. 81 84. Petroleum Geology 23, 5 53.
Eichhubl, P., Boles, J.R., 1998. Vein formation in relation to Jizba, D., Nur, A., 1990. Static and dynamic moduli of tight
burial diagenesis in the Miocene Monterrey Formation, gas sandstones and their relation to formation properties.
Arroyo Burro Beach, Santa Barbara, California. In: Eichhubl, Transactions of the SPWLA Annual Logging Symposium 31,
P. (Ed.), Diagenesis, Deformation, and Fluid Flow in the BB1 BB20.
Miocene Monterey Formation. Pacific Section SEPM, Long Lander, R.H., Walderhaug, O., 1999. Predicting porosity through
Beach, no. 83, pp. 15 36. simulating sandstone compaction and quartz cementation.
Engelder, T., 1985. Loading paths to joint propagation during a American Association of Petroleum Geologists Bulletin 83,
tectonic cycle: an example from the Appalachian Plateau, USA. 433 449.
Journal of Structural Geology 7, 459 476. Laubach, S.E., 1988. Subsurface fractures and their relationship to
Engelder, T., Geiser, P., 1980. On the use of regional joint sets as stress history in East Texas basin sandstone. Tectonophysics
trajectories of paleostress fields during the development of the 156, 37 49.
206 R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207

Laubach, S.E., 1989. Paleostress directions from the preferred Narr, W., Currie, J.B., 1982. Origin of fracture porosityexample
orientation of closed microfractures (fluid inclusion planes) in from the Altamont Field, Utah. American Association of
sandstone, East Texas basin, U.S.A. Journal of Structural Petroleum Geologists Bulletin 66, 1231 1247.
Geology 11, 603 611. Nelson, R.A., 1981. Significance of fracture sets associated with
Laubach, S.E., 1997. A method to detect natural fracture strike in stylolite zones. American Association of Petroleum Geologists
sandstones. American Association of Petroleum Geologists Bulletin 65, 2417 2425.
Bulletin 81, 604 623. Perez, R.J., Boles, J.R., 2004. Mineralization, fluid flow and sealing
Laubach, S.E., 2003. Practical approaches to identifying sealed and properties associated with a Quaternary thrust fault, San Joaquin
open fractures. American Association of Petroleum Geologists basin, California. American Association of Petroleum Geolo-
Bulletin 87 (4), 561 579. gists Bulletin 88 (9), 1295 1314.
Law, B.E., Spencer, C.W., 1998. Abnormal pressure in hydro- Perez, R.J., Gomez, M., Cassani, F., 1997. Integracion y diagenesis-
carbon environments. In: Law, B.E., Ulmishek, G.F., Slavin, petrofisica y calidad de reservorio en un area de Rio Pauji, Zulia
V.I. (Eds.), Abnormal Pressures in Hydrocarbon Environments. Oriental. 1er Congreso Latinoamericano de Sedimentologia,
Memoir-American Association of Petroleum Geologists, vol. 165 172.
70, pp. 1 11. Tulsa, OK. Perez, R.J., Chatellier, J.Y., Lander, R.H., 1999a. Use of quartz
Lugo, J., 1991. Cretaceous to Neogene tectonic control on cementation kinetic modeling to constrain burial histories.
sedimentation: Maracaibo Basin, Venezuela [unpublished PhD Examples from the Maracaibo basin, Venezuela. Revista Latin-
thesis]: University of Texas, Austin, Austin. oamericana de Geoquimica Organica 5, 39 46.
Lugo, J., Mann, P., 1995. JurassicEocene tectonic evolution Perez, R.J., Ghosh, S., Chatellier, J.Y., Lander, R.H., 1999b.
of Maracaibo Basin, Venezuela: Petroleum of South Application of sandstone diagenetic modeling to reservoir quality
America. Memoir-American Association of Petroleum Geolo- assessment of the Misoa Formation, Bachaquero Field, Mar-
gists, 699 725. acaibo basin, Venezuela. American Association of Petroleum
Lundegard, P.D., Land, L.S., 1986. Carbon dioxide and organic Geologists and SEPM Annual Meeting Abstract, p. A107.
acids: their role in porosity enhancement and cementation, Pittman, E., Larese, R., 1991. Compaction of lithic sands:
Paleogene of the Texas Gulf Coast. In: Gautier, D.L. (Ed.), experimental results and applications. The American Associa-
Roles of Organic Matter in Sediment Diagenesis. SEPM, Tulsa, tion of Petroleum Geologists Bulletin 75, 1279 1299.
pp. 129 146. Rodriguez, I., Navarro, A., Ghosh, S., 1997. Nueva frontera
Macaulay, C.I., Boyce, A.J., Fallick, A.E., Haszeldine, R.S., 1997. exploratoria en la Cuenca Petrolifera del Lago de Mar-
Quartz vein record vertical flow at a graben edge: Fulmar oil acaibo: Zulia Oriental, Venezuela Occidental. VI Simposio
field, central North Sea. American Association of Petroleum Bolivariano bExploracion petrolera en las Cuencas SubandinasQ,
Geologists Bulletin 81, 2024 2035. vol. 1, pp. 565 581.
Machel, H.G., 1987. Some aspects of diagenetic sulphatehydro- Roedder, E., 1984. Fluid inclusions. Mineralogical Society of
carbon redox reactions. In: Marshall, J.D. (Ed.), Diagenesis of America Reviews in Mineralogy 12, 644.
Sedimentary Sequences. Geological Society Special Publica- Roure, F., Colletta, B., De Toni, B., Loureiro, D., Passalacqua, H.,
tion, vol. 36, pp. 15 28. Liverpool, UK. Gou, Y., 1997. Within-plate deformations in the Maracaibo and
Maguregui, J., Tyler, N., 1991. Evolution of middle Eocene East Zulia basins, Western Venezuela. Marine and Petroleum
tide-dominated deltaic sandstones, Lagunillas Field, Mar- Geology 14, 139 163.
acaibo basin, western Venezuela. In: Miall, A.D., Tyler, Suppe, J., 1985. Principles of Structural Geology. Prentice-Hall,
N. (Eds.), Concepts in Sedimentology and Peleontology, Inc., Englewood Cliffs, New Jersey. 537 pp.
pp. 233 244. Surdam, R.C., Crossey, L.J., 1985. Mechanisms of organic/
Makowitz, A., Milliken, K.L., 2003. Quantification of brittle inorganic interactions in sandstone/shale sequences. In: Gautier,
deformation in burial compaction, Frio and Mount Simon D.L., Kharaka, Y.K., Surdam, R.C. (Eds.), Relationship of
Formation sandstones. Journal of Sedimentary Research 73, Organic Matter and Mineral Diagenesis. SEPM Short Course,
1007 1021. no. 17, pp. 177 232. Tulsa, OK.
Marret, R., Ortega, O., Kelsey, C., 1999. Extent of power- Swarbrick, R.E., Osborne, M.J., 1996. The nature and diversity of
law scaling for natural fractures in rock. Geology 27 (9), pressure transition zones. Petroleum Geoscience 2, 111 116.
799 802. Swarbrick, R.E., Osborne, M.J., 1998. Mechanisms that generate
Milliken, K.L., 1994. The widespread occurrence of healed micro- abnormal pressures: an overview. In: Law, B.E., Ulmishek, G.F.,
fractures in siliciclastic rocks: evidence from scanned cathodo- Slavin, V.I. (Eds.), Abnormal Pressures in Hydrocarbon
luminescence imaging. In: Nelson, P.P., Laubach, S.E. (Eds.), Environments. Memoir-American Association of Petroleum
Rock Mechanics. Balkema, Rotterdam, pp. 825 832. Geologists, pp. 13 34. Tulsa, OK.
Milliken, K.L., Laubach, S.E., 2000. Brittle deformation in Swarbrick, R.E., Osborne, M.J., Gareth, S.Y., 2002. Comparison of
sandstone diagenesis as revealed by scanned cathodolumi- overpressure magnitude resulting from the main generating
nescence imaging with application to characterization of mechanisms. In: Huffman, A., Bowers, G. (Eds.), Pressure
fractured reservoirs. In: Pagel, M., Barbin, V., Blanc, P., Regimes in Sedimentary Basins and their Prediction. Memoir-
Ohnenstetter, D. (Eds.), Cathodoluminescence in Geosciences. American Association of Petroleum Geologists, pp. 1 12.
Springer, pp. 225 243. Tulsa, OK.
R.J. Perez, J.R. Boles / Tectonophysics 400 (2005) 179207 207

Sweeney, J.J., Braun, R.L., Burnham, A.K., Talukdar, S., Vallejos, Vrolijk, P., Pottorf, R.J., Maze, W.B., 1996b. Record of source
C., 1995. Chemical kinetic model for hydrocarbon generation, generated overpressures, Venezuela. American Association of
expulsion, and destruction applied to the Maracaibo basin, Petroleum Geologists Bulletin 5, 145.
Venezuela. American Association of Petroleum Geologists Walderhaug, O., 1992. Magnitude of uplift of the Stb and Nordmela
Bulletin 79, 1515 1532. Formations in the Hammerfest Basina diagenetic approach.
Terzaghi, K., Peck, R.B., 1948. Soil Mechanics in Engineering Norsk Geologisk Tidsskrift 22, 321 333.
Practice. Wiley, New York, N.Y. 566 pp. Walderhaug, O., 1994. Precipitation rates for quartz cement in
Twiss, R.J., Moores, E.M., 1992. Structural Geology. W.H. Freeman sandstones determined by fluid-inclusion microthermometry
and Company, New York. 532 pp. and temperature-history modeling. Journal of Sedimentary
Van Veen, F.R., 1972. Ambientes sedimentarios de las Forma- Research, Section A 64, 324 333.
ciones Mirador y Misoa del Eoceno inferior y medio en la Walderhaug, O., 1996. Kinetic modeling of quartz cementation and
Cuenca del Lago de Maracaibo. Sedimentary environments of porosity loss in deeply buried sandstone reservoirs. American
the lower and middle Eocene Mirador and Misoa Formations Association of Petroleum Geologists Bulletin 80, 731 745.
of Lake Maracaibo Basin. Congreso Geologico Venezolano, Walderhaug, O., Bjbrkum, P.A., 2003. The effect of stylolite
1073 1104. spacing on quartz cementation in the Lower Jurassic Sto
Vasquez, Y., 1998. Caracterization Geoquimica De Las Aguas De Formation, southern Barents Sea. Journal of Sedimentary
Formacion Del Domo Sur, Campo Motatan, Cuenca Del Research 73, 145 156.
Lago De Maracaibo. Universidad Central de Venezuela-ICT. Warpinski, N.R., 1989. Elastic and viscoelastic calculations of
125 pp. stresses in sedimentary basins. SPE Formation Evaluation 4,
Vrolijk, P., Pottorf, R.J., Maze, W.B., 1996a. Overpressure 522 530.
history of fractures, west Maracaibo basin, Venezuela. Wood, S.A., Spera, F.J., 1984. Adiabatic decompression of aqueous
American Association of Petroleum Geologists Bulletin 80, fluid solutions: applications to hydrothermal fluid migration in
1344. the crust. Geology 12, 707 710.

Вам также может понравиться