Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Geochimica et Cosmochimica Acta, Vol. 62, No. 9, pp.

1599 –1617, 1998


Copyright © 1998 Elsevier Science Ltd
Pergamon Printed in the USA. All rights reserved
0016-7037/98 $19.00 1 .00
PII S0016-7037(98)00000-0

Laboratory and theoretical constraints on the generation and composition of natural gas
JEFFREY S. SEEWALD,* BRYAN C. BENITEZ-NELSON, and JEAN K. WHELAN
Department of Marine Chemistry and Geochemistry, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts 02543, USA

(Received September 12, 1996; accepted in revised form January 20, 1998)

Abstract—Hydrous pyrolysis experiments were conducted at 125 to 375°C and 350 bars to constrain factors
that regulate the generation and relative abundance of hydrocarbon and nonhydrocarbon gases during thermal
maturation of Monterey, Eutaw, and Smackover shale. Thermogenic gas was generated at temperatures as low
as 125°C and increased in abundance with increasing temperature. The relative abundance of individual
hydrocarbons varied substantially in response to increasing time and temperature reflecting the chemical
processes responsible for their formation. The hydrocarbon fraction of low maturity gas produced via primary
cracking of kerogen was composed predominantly of methane. With increasing thermal maturity, the onset of
bitumen generation produced longer-chain hydrocarbons causing a decrease in the relative abundance of
methane. At high levels of thermal maturity, the absolute and relative abundance of methane increased due to
decomposition of bitumen.
In all experiments at all temperatures, carbon dioxide was the most abundant volatile organic alteration
product. Carbon dioxide was produced directly from kerogen at low thermal maturity and via the decompo-
sition of bitumen and/or kerogen at high thermal maturity. During early stage alteration, kerogen likely
represents the dominant source of oxygen in carbon dioxide while at high thermal maturities water may
represent an abundant and reactive oxygen source. Hydrogen released during the disproportionation of water
is likely consumed during hydrocarbon generation. Theoretical reaction path modeling suggests that the
precipitation of calcite may effectively remove carbon dioxide from natural gas if a source of Ca is available
within the rock. Thus, carbon dioxide-rich natural gas may be relatively pristine while methane-rich natural
gas may reflect the occurrence of secondary reactions involving inorganic sedimentary components.
Kinetic analysis of the experimental data indicates a narrow range of activation energies for the generation
of C1-C4 hydrocarbons from the Monterey, Smackover, and Eutaw shales. Carbon dioxide generation from the
Monterey and Eutaw shales is accounted for by a substantially broader range of activation energies.
Application of these data to predict gas formation at temperatures and time scales typical of subsiding
sedimentary basins suggests that C1-C4 generation is restricted to relatively high temperatures while carbon
dioxide generation occurs at both low and high thermal maturities. Thus, in contrast to the bulk of C1-C4
generation which is predicted to occur after peak bitumen generation, production of carbon dioxide will occur
before, during, and after the generation of liquid hydrocarbons. Copyright © 1998 Elsevier Science Ltd

1. INTRODUCTION products are often obscured. Laboratory heating experiments


are an effective tool for constraining organic alteration pro-
Thermal maturation of sedimentary organic matter involves
cesses under well-constrained physical and chemical condi-
numerous chemical transformations that result in the produc-
tions. The generation of oil and natural gas has been studied
tion of volatile hydrocarbon and nonhydrocarbon alteration
using a variety of laboratory techniques including open-system
products. In addition to forming economic deposits, thermo-
dry pyrolysis (Espitalié et al., 1977; Burnham et al., 1987),
genic gases influence many geologic processes associated with
closed-system dry pyrolysis (Monthioux et al., 1985; Schenk
the transport and accumulation of petroleum in sedimentary
and Horsfield, 1993), and closed-system hydrous pyrolysis
basins. In particular, carbon dioxide, a quantitatively important
component of natural gas, not only represents a major contam- (Lewan, 1993a). The generation rates and composition of or-
inant from an economic perspective, but also influences reser- ganic alteration products are highly dependent on the simula-
voir porosity and permeability during sediment diagenesis due tion technique employed. In particular, hydrous pyrolysis ex-
to its weak acid properties and the ability to form carbonate periments have demonstrated that in addition to influencing the
minerals. Understanding fundamental processes that regulate amounts and composition of liquid hydrocarbons, the presence
the generation of organically derived gaseous products is there- of liquid water plays an important role in the generation of
fore essential to constrain factors influencing the accumulation volatile organic alteration products (Lewan, 1993b, 1997). The
of oil and gas in sedimentary basins. reasons for this are numerous and likely include changes in
Field studies have provided us with a wealth of data regard- reaction mechanism and the possible role of water as a source
ing the occurrence and composition of natural gas. Because of hydrogen and oxygen for the formation of organic alteration
natural systems are inherently complex, processes that regulate products such as methane and carbon dioxide (Hoering, 1984;
the timing, amounts, and composition of organic alteration Siskin and Katritzky, 1991; Helgeson et al., 1993; Lewan,
1993a,b, 1997; Price, 1993; Seewald, 1994, 1996; Stalker et al.,
1994). Because, liquid water is an abundant phase in almost all
*Author to whom correspondence should be addressed (jseewald@ sedimentary basins and is generally present in petroleum res-
whoi.edu). ervoirs during the conversion of oil to natural gas, it is essential
1599
1600 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

that the role of water during the thermal maturation of sedi-


mentary organic matter to produce volatile alteration products
such as methane and carbon dioxide be considered.
Except for a few notable exceptions (Andresen et al., 1994;
Lewan, 1997), hydrous pyrolysis studies have focused on con-
straining the generation of hydrocarbons from kerogen and
have provided only limited data for the generation of organi-
cally derived carbon dioxide. Quantification of carbon dioxide
generation during hydrous pyrolysis is complicated by the fact
that multiple sources and sinks for carbon dioxide may exist.
For example, determination of organically derived carbon di-
oxide must account for dissolution of inorganic carbonate min-
erals during an experiment. Alternatively, carbon dioxide may
precipitate as carbonate minerals if sufficiently high dissolved
concentrations are attained. Moreover, due to the high solubil-
ity of carbon dioxide in water, underestimates of carbon diox-
ide production may be obtained from chemical analysis of the
gas phase alone during hydrous pyrolysis (Lewan, 1997). Ac-
cordingly, experiments conducted during this study were de-
signed to circumvent these problems and provide data con-
straining the generation of organically derived carbon dioxide
and low molecular weight volatile hydrocarbons as a function
of time and temperature. These data are used to develop chem-
ical and kinetic models that account for the generation and
composition of thermogenic gases in sedimentary basins.

2. METHODS
2.1. Sample Description and Preparation
Samples of Monterey, Eutaw, and Smackover shales were utilized
for this study (Table 1). The Monterey Shale is a Miocene sediment
collected at an outcrop in Naples Beach, California, USA. The sample
is an unweathered lenticularly laminated organic-rich (20.6 wt% TOC)
claystone that is thermally immature (bitumenite reflectance 5 0.25%
expressed as Ro). The Eutaw Shale is a Cretaceous sediment obtained
from core cuttings recovered at a depth interval of 3011 to 3021 m from
the Hampton/Nelson Ball well in Pike County, Mississippi, USA. It is
a quartz-rich, organic lean shale (0.97 wt% TOC) with measured
vitrinite reflectance of 0.45% Ro, a hydrogen index of 60, and a Tmax
of 428°C. The Smackover Shale is a Jurassic sediment obtained from
core cuttings recovered at a depth of 3260 m from the Amareda Scotch
Well #1 in Clark County , Alabama, USA. This shale is also quartz-rich
and organic-lean (0.92 wt% TOC) with a measured vitrinite reflectance
of 0.52% Ro, a hydrogen index of 67, and a Tmax of 439°C. Based on
the hydrogen index, Tmax values, and petrographic observations, or- not produce detectable amounts of low molecular weight hydrocarbons
ganic matter in the Eutaw and Smackover shales is predominantly of indicating the effective removal of sorbed gases.
terrigenous origin that corresponding to a type-III kerogen (Espitalié et
al., 1984). Elemental analysis indicates that the Monterey Shale con- 2.2. Experimental Apparatus and Conditions
tains type II-S kerogen (Table 1).
In this study our primary goal was to constrain the rate and extent of Except for experiment MS15, all experiments were conducted in 316
chemical reactions responsible for the generation of gaseous products. stainless steel tubing reactors with a 20 cm3 internal volume (Fig. 1).
Accordingly, to minimize physical processes associated with expulsion The tubing reactor permits fluid samples to be withdrawn from the
of produced gases, the starting rocks were ground in a disc mill until the pipe-bomb at the temperature and pressure of an experiment allowing
entire sample passed through a ,125 mm sieve. To remove existing fluid chemistry to be monitored as a function of time. The experiments
hydrocarbons, ground samples of the Eutaw and Smackover shales described here are in many ways similar to the hydrous pyrolysis
were Soxhlet extracted in a 93:7 dichloromethane/methanol mixture for experiments described by Lewan (1993a) in that bulk rocks are heated
48 h while the Monterey Shale was sequentially extracted by sonic in the presence of excess liquid water, but there are important differ-
disruption for 9 min in methanol, a 1:1 mixture of dichloromethane/ ences. For example, in our experiments the reactor vessel is filled
methanol, and dichloromethane. Except for experiments MS1 and completely with liquid water precluding the existence of a vapor
MS2, all starting ground shale samples were treated with 10% hydro- headspace. In addition, to minimize the possibility that the amounts of
chloric acid at 40°C for 2 h following solvent extraction to remove generated gases would exceed their aqueous solubility and form a
inorganic carbonate minerals. Coulometric analysis for inorganic car- separate phase, only 1.0 g of bulk rock was used for each experiment
bon following this treatment revealed carbonate was completely re- resulting in fluid/rock mass ratios considerably higher than those typ-
moved. Thus, carbon dioxide generated during the experiments is ically associated with conventional hydrous pyrolysis. The measured
organic in origin and not the result of carbonate mineral dissolution. concentrations of gases remained below their aqueous solubilities at the
Sonication of the ground, solvent-extracted, and acid-treated rocks in conditions of all experiments (Bonham, 1978; Price, 1981; Bowers and
water at 90°C for 15 min while purging with a helium carrier gas did Helgeson, 1983; Johnson et al., 1992). Although the presence or
Generation and formation of natural gas 1601

pumping Ar-purged distilled water into one end of the tubing reactor
while the fluid sample was removed from the opposite end. The
sampling process was performed rapidly (1 to 2 min) to avoid dilution
of the fluid samples with the freshly injected water. The absolute
amounts of generated gases in subsequent samples were corrected for
gases removed during earlier sampling events.
Experiment MS15 differed substantially from those described above
and was intended to preclude precipitation of carbonate minerals that
could act as a sink for organically derived carbon dioxide. Instead of
containing deionized water as the starting fluid, this experiment con-
tained 55 g of a Na-Mg-Cl fluid of seawater salinity (Table 3). In
addition, 1.52 g of Monterey Shale and 0.51 g of finely ground pure
quartz were added to the reaction cell. The Mg-rich fluid and quartz
result in the maintenance of low in-situ pH conditions as a result of talc
precipitation according to the reaction:

3Mg 11 1 4SiO 2~qtz)14H2O(l)5Mg3Si4O10(OH)2(s)16H1 (1)

(Saccocia and Seyfried, 1991). By maintaining low in-situ pH, the


solubility of calcite is considerably enhanced according to the reaction:

Ca 11 1 H 2CO 3~aq)5CaCO3(s)12H1 (2)

relative to the deionized water experiments. Experiment MS15 was


conducted at 225 to 375°C and 350 bars in a similar fashion to the
Eutaw Shale experiment in that temperature was maintained at a
constant value before incremental increases at selected time intervals.
Because the critical point of the saline fluid used in experiment MS15
is 407°C (Bischoff and Rosenbauer, 1985) this experiment was con-
ducted under subcritical conditions.
Owing to the low pH and corrosiveness of the high chlorinity fluids
in experiment MS15, a chemically inert Au and Ti flexible-cell hydro-
thermal apparatus was used (Seyfried et al., 1987). This equipment also
allows external control of pressure ensuring a single liquid phase inside
the reaction cell. An important difference compared to the tubing
reactor experiments, however, is that water is pumped into the pressure
vessel volume surrounding the reaction cell to maintain pressure during
sampling. Accordingly, the reaction cell fluid is not diluted during
sampling. However, as result of fluid sampling, the fluid/rock mass
ratio decreases during the experiment.
Fig. 1. Tubing reactor utilized for this study.

2.3. Analytical Procedures


absence of an oil phase could not be determined during an experiment,
the high fluid/rock mass ratios employed minimized the potential for Fluid samples were analyzed for the dissolved concentra-
the formation of such a phase due to the relatively high solubility of oil tions of carbon dioxide, C1-C4 hydrocarbons, H2, and H2S.
in water at the temperatures of these experiments (Price, 1981), espe- Dissolved carbon dioxide and C1-C4 hydrocarbons were ana-
cially for the organic-lean Smackover and Eutaw shales. In each lyzed by injecting fluid samples into a purge and trap apparatus
experiment it was assumed that all gases were present as dissolved
species in the aqueous phase allowing determination of the absolute
interfaced directly to a gas chromatograph equipped with flame
amounts generated from analyses of the aqueous phase alone. The ionization and thermal conductivity detectors. The fluid sam-
absence of a vapor headspace in our experiments is important since ples were acidified with 1 mL of 25 % phosphoric acid in the
reactions are limited to those involving aqueous species and solid purge cell to ensure complete evolution of dissolved carbon
phases. In contrast, conventional hydrous pyrolysis may involve reac- dioxide. Separation of gases in the gas chromatograph was
tions occurring in oil, water, gas, and solid phases.
Prior to heating, the tubing reactor was evacuated with a vacuum achieved with either a porapak-Q packed column or a po-
pump and partially filled with Ar-purged distilled water to ensure that raplot-Q megabore capillary column. Following a headspace
the rock was wet during the 30 min heat-up process. Once at the desired extraction, dissolved H2 concentrations were determined in
temperature, the tubing reactor was completely filled with water and selected experiments using a gas chromatograph equipped with
pressurized to 350 bars. Because the specific volume of water increases
with increasing temperature at constant pressure, the fluid/rock mass
a thermal conductivity detector and a 5 Å molecular sieve
ratio was lower in the high-temperature experiments. Temperature was packed column. Dissolved H2S was determined gravimetrically
monitored (62°C) with a thermocouple at each end of the tubing by acidifying the fluid sample with 25 % phosphoric acid and
reactor to ensure there were no thermal gradients. Experiments utilizing precipitating the evolved H2S as Ag2S in a 3 wt% AgNO3
the Monterey and Smackover shales were conducted isothermally for solution. The concentrations of dissolved Na, Ca, K, Mg, and
approximately 170 h. In contrast, the Eutaw Shale investigation was
conducted by maintaining the tubing reactor at a constant temperature Cl in experiment MS15 were analyzed by ion chromatography.
for a selected period of time before rapidly increasing the temperature Careful accounting of all fluid removed and added to the
to a new value. At each temperature condition gas generation was tubing reactor allowed the absolute amounts of generated prod-
monitored as a function of time. ucts to be calculated from the dissolved concentrations, tubing
During the course of an experiment, fluid samples were withdrawn
into glass gas-tight syringes and analyzed for the concentration of
reactor volume, and the mass of rock in each experiment. The
dissolved gases. An internal 10 mm steel mesh filter ensured particle specific volumes of water at the temperature and pressure of
free fluids. Pressure was maintained during the sampling process by each experiment were calculated using the equation of state of
1602 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan
Generation and formation of natural gas 1603

associated with the choice of standards. This is especially true


for analysis of bitumen containing abundant asphaltenes since
FID response is strongly dependent upon asphaltene composi-
tion. Accordingly, uncertainties associated with these data are
estimated at approximately 30 %. Insufficient bitumen extract
from the Eutaw and Smackover shale experiments precluded
this analysis.
Elemental analyses (CHNSO) were performed on all starting
rocks and solid alteration product. Inorganic carbonate was
determined on a Coulometrics carbon analyzer. Because inor-
ganic carbon abundances were below detection (,0.03 wt%) in
all starting rocks and alteration products, total organic carbon
(TOC) is equivalent to the total carbon determined by elemen-
tal analysis. The starting rock and solid alteration products for
the Monterey Shale experiments were demineralized by HCl
and HF acid digestions (Eglinton and Douglas, 1988) to isolate
kerogen. The isolated kerogens from all experiments except
MS15 were treated with CrCl2 and HCl to remove iron sulfides
(Canfield et al., 1986; Acholla and Orr, 1993) before being
analyzed for their elemental composition. Similar kerogen iso-
Haar et al. (1980). Analytical uncertainties are estimated to be lations for the Eutaw and Smackover shales were not possible
,5 % for all dissolved species. owing to the low organic matter content of these rocks.
After completion of each experiment the liquid reactants
were removed by sequentially pumping 30 mL of methanol and 3. RESULTS
dichloromethane through the tubing reactor at room tempera- 3.1. Volatile Products and Generated Bitumen
ture. The solid reactants were then removed from the tubing
reactor, dried in air at 40°C, and solvent extracted by sonic Thermal maturation of organic-bearing sedimentary rocks is
disruption as described above. This extract was combined with characterized by the generation of gaseous alteration products,
the tubing reactor rinses and back-extracted with distilled H2O. the extent and rate of which are strongly dependent on time and
The combined extract, referred to here as total bitumen, in- temperature (Tables 2–5; Figs. 2–5). For example, the amount
cludes expelled oil and differs from bitumen measured during of dissolved methane generated during heating of Monterey
conventional hydrous pyrolysis which is limited to products Shale for 170 h increased from 0.0036 mmole/g TOC at 125°C
retained in the source rock (Lewan, 1993a). Quantification of to 2.9 mmole/g TOC at 375°C. Continued reaction at 375°C
the total bitumen extract from the Monterey Shale experiments resulted in the eventual generation of 5.8 mmole/g TOC meth-
was performed by thin-layer chromatography with flame ion- ane after 1875 h. Production of methane during heating of the
ization detection (TLC-FID) according to the methods of Smackover and Eutaw shales showed a similar temperature
Karlsen and Larter (1991). The TLC-FID technique is not dependence, although at a given temperature and reaction time
particularly precise and accuracy suffers due to uncertainties the absolute amounts of methane were considerably less than
1604 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

Fig. 2. Amounts of low molecular weight organically derived alteration products generated as a function of time and
temperature during hydrous pyrolysis of Monterey Shale for 170 h (symbols). The dashed lines indicate predicted gas
generation during the experiments using a parallel reaction model and kinetic parameters retrieved from the experimental
data (see text).

the Monterey Shale. The absolute abundance of ethane, pro-


pane, and butane also increased with increasing temperature,
but remained below that of methane in all experiments.
Variations in the absolute abundance of C1-C4 hydrocarbons
were accompanied by large changes in their relative abundance
in the hydrocarbon fraction of generated gas (Fig. 6). During
heating of Monterey Shale, methane abundance constituted
approximately 90 mole % of the C1-C4 hydrocarbons at 125 to
200°C before systematically decreasing to values as low as 58
mole % at 325°C. At temperatures .325°C, the relative abun-
dance of methane increased to a final value of 68 mole % after
1875 h of reaction at 375°C. Almost identical variations in the
relative abundance of methane with respect to C1-C4 hydrocar-
bons were observed during the Eutaw Shale experiment except
that the initial decrease in wt% methane did not occur until a
temperature of 325°C. Because bitumen in the initial shales
used for these experiments was removed by solvent extraction
Fig. 3. Amounts of low molecular weight organically derived alter- prior to heating, generation of methane-rich gas at low temper-
ation products generated as a function of time and temperature during atures prior to the onset of bitumen generation is consistent
Monterey Shale experiment MS15 (symbols). The dashed lines indicate
predicted gas generation during the experiments using a parallel reac-
with derivation via the primary cracking of kerogen. Subse-
tion model and kinetic parameters retrieved from the experimental data quent decreases in mole % methane are likely associated with
(see text). the generation of bitumen containing high molecular weight
Generation and formation of natural gas 1605

Fig. 4. Amounts of low molecular weight organically derived alteration products generated as a function of time and
temperature during hydrous pyrolysis of Smackover Shale for 170 h (symbols). The dashed lines indicate predicted gas
generation during the experiments using a parallel reaction model and kinetic parameters retrieved from the experimental
data (see text).

compounds. Indeed, decreases in the wt% methane during the the decrease in wt% methane at 325°C suggests higher tem-
Monterey Shale experiments coincided with the onset of bitu- peratures may be required for peak bitumen generation. The
men generation which increased substantially at 200°C and rate of petroleum generation has been shown to correlate pos-
peaked at 275°C (Fig. 7). Although bitumen generation was not itively with kerogen S content (Hunt et al., 1991). Thus, lower
quantitatively monitored during the Eutaw Shale experiments, temperatures for bitumen generation from the Monterey Shale
are consistent with its relatively high S content.
In contrast to the Monterey and Eutaw shales, gas generation
during the low temperature Smackover Shale experiments was
characterized by lower mole % methane in the hydrocarbon
fraction (Fig. 6). With increased temperature, however, the
relative abundance of methane increased continuously reaching
a value of 82 mole % at 360°C. The absence of a methane-rich
gas in the low-temperature Smackover Shale experiments may
reflect a compositional difference in the kerogen due to varia-
tions in the sources of organic matter in these sediments and/or
variations in their digenetic histories prior to collection.
The relative abundance of methane in the hydrocarbon frac-
tion increased continuously during maturation at temperatures
.325°C in all experiments indicating the production of meth-
ane-rich gas at high thermal maturity. The progression with
increasing temperature from a system generating predomi-
Fig. 5. Amounts of low molecular weight organically derived alter- nantly bitumen to one producing gas is also reflected in the
ation products generated as a function of time and temperature during sharp increase in the gas/bitumen volume ratio at 275°C (Fig.
hydrous pyrolysis of Eutaw Shale (symbols). The dashed lines indicate
7). Decreases in the TOC content of the Monterey Shale at
predicted gas generation during the experiments using a parallel reac-
tion model and kinetic parameters retrieved from the experimental data temperatures .325°C account for only a minor fraction of the
(see text). carbon contained in the generated gas suggesting decomposi-
1606 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

Fig. 6. Relative abundance of organically derived methane and carbon dioxide as a function of temperature and/or time
hydrous pyrolysis of (a) Monterey Shale, (b) Smackover Shale, and (c) Eutaw Shale. All data points for the Monterey and
Smackover Shale represent gas composition after 170 h of heating except for the 1875 h labeled Monterey Shale data
connected by dashed lines.

tion of bitumen is primarily responsible for generation of low alteration products from the Monterey Shale experiments is
molecular weight hydrocarbons. The decrease in the abundance provided in Nelson et al. (1995).
of bitumen at temperatures .275°C during the Monterey Shale Carbon dioxide was unequivocally the dominant gaseous
experiments (Fig. 7) supports such a model. This decrease alteration product released to aqueous solution during all ex-
primarily reflects lower abundances of polar material (Nelson periments. Because inorganic carbonate was removed from the
et al., 1995) indicating production of the low molecular weight starting rock samples prior to each experiment, the generated
species via the decomposition of asphaltenes and resins. A carbon dioxide was entirely organic in origin. In general, the
more detailed discussion of the bitumen extracts and solid amounts of carbon dioxide released to solution increased with
Generation and formation of natural gas 1607

precluded carbonate mineral precipitation resulting in system-


atic increases in dissolved carbon dioxide as a function of time
and temperature. The measured concentrations at 325°C were
not substantially greater during experiment MS15 relative to
MS5 and MS13 suggesting that carbonate precipitation was not
extensive during the latter two experiments.
During experiments MS1 and MS2 in which inorganic car-
bonate was not removed from the starting sediment, measured
carbon dioxide abundances were less than or the same as in
experiments conducted at lower temperatures (Fig. 2). This
result is opposite to what one might predict considering the
potential for carbonate minerals initially present in the sedi-
ment to dissolve. In the presence of carbonate minerals, how-
ever, the abundance of carbon dioxide in solution will be
regulated by the solubility of minerals such as calcite (CaCO3)
and siderite (FeCO3), provided kinetic constraints do not pre-
clude equilibration with the fluid. Apparently, during these
experiments, pH and the activities of dissolved Ca11 and/or
Fe11 resulted in lower dissolved carbon dioxide abundances.
In addition to saturated hydrocarbons, significant quantities
of alkenes were generated at all temperatures for all rocks. In
general, the abundance of alkenes increased with increasing
temperature. In contrast to the saturated hydrocarbons, how-
ever, the abundance of alkenes decreased with continued reac-
tion at constant temperature, except during the 125°C experi-
ments in which their abundance remained constant. These
results indicate that ethene, propene, and n-butene are unstable
at the concentrations generated during these experiments at
Fig. 7. Gas/bitumen volume ratio (a) and amount of total bitumen
temperatures $150°C. That the alkene abundance at 125°C
generated (b) as a function of temperature during hydrous pyrolysis of remained constant while the abundance of alkanes increased
Monterey Shale for 170 h. may suggest that the alkenes were being generated but degra-
dation reactions were occurring at a similar rate. These data are
consistent with the results of Tannenbaum and Kaplan (1985)
increasing time and temperature and are consistent with results who suggested that alkenes are initially produced in small
from previous hydrous pyrolysis experiments (Cooles et al., quantities under dry or hydrous conditions but represent an
1987; Burnham et al., 1992; Barth and Nielsen, 1993; Andresen unstable intermediate during the formation of petroleum.
et al., 1994, 1995). Some of the experiments, however, were Total dissolved H2S concentrations during the Monterey and
characterized by carbon dioxide abundances that demonstrated Eutaw shale experiments generally increased with increasing
nonsystematic variations with respect to time and temperature. temperature. At temperatures #225°C dissolved H2S concen-
For example, heating of the Smackover Shale at 250 and 325°C trations were too low for gravimetric determination, but H2S
produced almost identical carbon dioxide abundances while was detectable by odor in all experiments. The H2S concentra-
heating of Monterey Shale at 325°C during experiments MS5 tion was not determined during the Smackover Shale experi-
and MS13 produced carbon dioxide abundances that were ments but a strong H2S odor in all fluid samples indicated its
almost identical to experiment MS11 at 300°C. These nonsys- presence. The absolute abundance of total dissolved H2S was
tematic variations were not reflected in the abundances of likely regulated by the solubility of iron sulfides such as pyrite
hydrocarbon gases measured in the same samples suggesting and/or pyrrhotite owing to the rapid dissolution-precipitation
carbonate mineral precipitation may be responsible for removal kinetics for these minerals under hydrothermal conditions (See-
of carbon dioxide from solution. Relatively erratic trends in the wald and Seyfried, 1990). Pyrite was abundant in the unheated
abundance of carbon dioxide as a function of time and temper- Monterey Shale and persisted along with newly formed pyr-
ature or cases where carbon dioxide decreases with increasing rhotite in the thermally altered rocks. Sources of S during these
temperature are common during hydrous pyrolysis (Andresen experiments include diagenetic pyrite and organic matter. Al-
et al., 1994; Burnham et al., 1992; Cooles et al., 1987) and though it is not possible to quantitatively determine the relative
suggests that precipitation of carbonate minerals such as calcite contributions from these two sources, the amounts of H2S
and/or siderite may be a common phenomenon. released to solution at 350, 360, and 375°C for the Monterey
There was no indication of carbonate mineral precipitation Shale are 13, 19 and 27 mg/g rock, respectively. These values
during experiment MS15. Changes in fluid composition during exceed the amount of S present as diagenetic pyrite in the
this experiment are consistent with the generation of acidity starting rock (10.5 mg/g rock as H2S) and provide evidence for
according to reaction 1. Dissolved Mg concentrations de- the release of organically bound S to solution.
creased at temperatures .275°C and measured pH (25°C) Dissolved H2 concentrations were determined in selected
varied from 2.91 to 3.52 (Table 3). Accordingly, low pH likely fluid samples from the Monterey Shale experiments and in
1608 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

bon sediment components. Mass balance calculations indicate


that this increase requires dissolution of approximately 15 wt%
of the inorganic material present.

4. DISCUSSION

Conventional models that account for the timing and com-


position of thermogenic organic alteration products are to a
large extent based on compositional trends for kerogen that
occur in response to increased maturation (Tissot and Welte,
1984; Hunt, 1996). In general, early stage kerogen maturation
is characterized by large decreases in the O/C ratio due to loss
of O-bearing carboxyl, carbonyl, and methoxy functional
groups, with only minor decreases in the H/C ratio (Fig. 8).
With continued burial and heating the O/C ratio remains con-
stant as the supply of labile oxygen is exhausted, while the
onset of hydrocarbon generation results in large decreases in
the H/C ratio. Because it is generally assumed that kerogen
represents the only source of oxygen and hydrogen available
for incorporation in organic alteration products, production of
O-bearing compounds such as carbon dioxide is generally
viewed as being limited to the early phases of kerogen matu-
ration prior to peak oil generation (Tissot and Welte, 1984;
Ungerer, 1990; Cooles et al., 1986; Barker, 1990).
Variations in the compositional evolution of kerogen during
the Monterey Shale experiments were generally consistent with
maturation trends in nature (Table 1; Fig. 8). Large decreases in
the kerogen O/C ratio were observed in the relatively low-
Fig. 8. Van Krevelen diagram showing compositional evolution of temperature experiments with only a minor decrease in the H/C
kerogen with increasing thermal maturity (adapted from Hunt, 1996). ratio, while maturation at relatively high temperatures resulted
The solid circles show changes in kerogen composition during hydrous
in only minor decreases in O/C ratio and substantially greater
pyrolysis of Monterey Shale.
decreases in the H/C ratio. The results of this study, however,
suggest that changes in kerogen composition may not accu-
samples from the 325 and 360°C phases of the Eutaw Shale rately reflect the timing of carbon dioxide generation. Al-
experiment. During the Monterey Shale experiments the dis- though, significant quantities of carbon dioxide were generated
solved H2 concentrations showed a systematic increase with during rock maturation at low temperatures, carbon dioxide
increasing temperature, except for experiment MS5 at 325°C. continued to be a major organic matter alteration product in the
This experiment was the first experiment conducted in a new experiments at high temperatures. Thus, experimental evidence
tubing reactor and the relatively high dissolved H2 concentra- suggests that organically derived carbon dioxide was generated
tions may reflect disproportionation of water during oxidation before, during, and after peak hydrocarbon generation.
of the stainless steel tubing reactor. Similarly, the Eutaw Shale Comparison of results for carbon dioxide evolution during
experiment was also conducted in a new tubing reactor and the hydrous and dry pyrolysis reveals a critical role for water
relatively high dissolved H2 concentrations at 325°C may re- during organic matter maturation. Heating of isolated Monterey
flect oxidation of the reactor walls. Continued reaction at con- Shale kerogen during closed-system dry experiments resulted
stant temperature resulted in decreasing dissolved H2 concen- in a maximum carbon dioxide yield of approximately 2.3
trations. mmole/g TOC after heating at 375°C for 24 h with no further
increases at temperatures as high as 500°C (Tomic et al., 1995).
This amount of carbon dioxide was observed during our ex-
3.2. Solid Alteration Products
periments after heating at only 250°C for 24 h and is consid-
In general, the bulk sediments were characterized by de- erably lower than the maximum carbon dioxide production of
creases in TOC, consistent with the generation of hydrocarbons 8.0 mmole/g TOC at 375°C. The difference in carbon dioxide
(Table 1). The N and S content of the Monterey bulk sediment yields is greater than can be accounted for by differences in the
also decreased. Analyses of the Monterey Shale isolated kero- oxygen content of the initial kerogens used for each study and
gens reveal increasing C and decreasing O and H contents with suggests that water may be contributing oxygen for carbon
increasing temperature. dioxide production in our experiments. Thus, greater yields and
Examination of Table 1 reveals a minor but significant initial generation at higher thermal stress suggest additional and/or
increase in the TOC content of the solid products during low different reactions are responsible for the production of carbon
temperature maturation of the Monterey Shale. Because the dioxide in the presence of water relative to dry systems.
sediment represents the only source of C present during the Mass balance calculations can be used to constrain sources
experiments, this increase likely reflects dissolution of noncar- of oxygen for carbon dioxide production. These calculations
Generation and formation of natural gas 1609

Laboratory experiments have documented the reaction of


individual hydrocarbons with water to produce oxygenated
alteration products such as ketones and alcohols at temper-
ature conditions similar to those employed during this study
(Leif and Simoneit, 1995; Seewald, 1996). These com-
pounds decompose to form carbon dioxide with continued
heating (Seewald, 1996). Involvement of water and carbon
dioxide during maturation of petroleum was suggested by
Helgeson et al. (1993) after a theoretical evaluation of
petroleum compositions. Thus, in the presence of water,
decomposition of petroleum at elevated temperatures and
pressures may not be limited to the production of short-chain
hydrocarbons and a C-rich residual as invoked by traditional
models (Tissot and Welte, 1984; Hunt, 1996), but may
involve the production of significant quantities of carbon
dioxide.
Utilization of water-derived oxygen during the production of
O-rich alteration products would necessarily result in the re-
lease of hydrogen. Formation of saturated hydrocarbons via the
termination of alkyl radicals or the saturation of alkenes to form
alkanes is likely responsible for the consumption of water-
derived hydrogen during maturation of organic matter. The
Fig. 9. Plot of oxygen released from kerogen (white bars) and incorporation of water-derived hydrogen in hydrocarbons is a
oxygen measured in solution as carbon dioxide during hydrous pyrol- facile process at elevated temperatures (Hoering, 1984; Stalker
ysis of Monterey Shale (black bars) as a function of temperature.
et al., 1994). In particular, results of redox-buffered hydrother-
mal experiments have demonstrated the reaction of alkenes
indicate that changes in the oxygen content of kerogen during with water to form alkanes (Seewald, 1994, 1996). Alkenes, a
the experiments are sufficient to account for the quantities of known product during the pyrolysis of sedimentary organic
carbon dioxide generated during heating of Monterey Shale at matter under dry and aqueous conditions (Tannenbaum and
125 to 360°C (Fig 9). At 375°C, however, the amount of Kaplan, 1985; Esser and Schwochau, 1991), were present in the
oxygen as carbon dioxide exceeded that released from the early stages of these experiments and showed decreasing abun-
kerogen by 7.4 % after 1875 h. This oxygen deficit represents dance with increasing time, even at temperatures as low as
a minimum value, however, since the generation of other 150°C. This trend is consistent with the production of alkenes
O-bearing alteration products such as water, carboxylic acids, as an unstable reaction intermediary during the thermal matu-
and bitumen were not considered in these calculations. In ration of kerogen and bitumen, and subsequent saturation to
particular, organically derived water has been shown to con- form alkanes.
sume approximately 30 % of organic oxygen during confined The possibility that water may represent a source of oxygen
pyrolysis of type II and III kerogens (Monthioux et al., 1985; and hydrogen for the formation of carbon dioxide and hydro-
Monthioux, 1988; Behar et al., 1992). carbons has important implications for the generation of natural
Additional sources of oxygen during the experiments include gas in sedimentary basins. Because water is ubiquitous in most
added water and inorganic rock components. Considering the sedimentary basins, predictive models may underestimate the
low ferric iron content of the Monterey Shale and thermody- gas potential of sedimentary organic matter if they assume
namic barriers to reducing ferrous iron and aluminosilicate hydrogen and oxygen in organic alteration products are derived
minerals, oxygen derived directly from these sources is not from kerogen only. It is unlikely that hydrogen availability
available for the formation of O-bearing organic alteration represents the limiting factor during the initial generation of
products. Accordingly, water is indicated as an important liquid hydrocarbons because cleavage of relatively intact kero-
source of oxygen for the formation of carbon dioxide during gen fragments to produce high molecular weight saturated
organic matter maturation. This conclusion is consistent with alteration products requires only minor amounts of hydrogen.
the results of other hydrous pyrolysis studies that have impli- With increasing thermal maturation, however, hydrogen de-
cated water as a source of oxygen in organic alteration products mands increase as organic transformations result in a decrease
at temperatures $300°C (Stalker et al., 1994; Lewan, 1997). in the average carbon chain length of hydrocarbons and asso-
Carbon dioxide generation at high temperature may occur ciated increases in the H/C ratio. Thus, inclusion of water as a
directly from kerogen or via the decomposition of bitumen. reactant during organic transformations will greatly increase
Because heating of Monterey Shale at temperatures .325°C estimates for the amount of carbon dioxide and hydrocarbon
was not accompanied by decreases in the oxygen content of gas that can be generated deep within a sedimentary basin.
kerogen (Fig. 9), production of carbon dioxide directly from Comparison of compositional trends observed during the
kerogen at these temperatures requires incorporation of wa- experiments with observations from nature show both similar-
ter-derived oxygen. Requisite oxygen for the generation of ities and differences. The evolution of the hydrocarbon fraction
carbon dioxide during the decomposition of bitumen could of natural gas from being almost entirely composed of methane
be derived from oxygen present in bitumen and/or water. at shallow depths to containing higher abundances of C2-C7
1610 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

hydrocarbons at greater sediment depths has been well docu- bons during the decomposition of bitumen in the presence of
mented (Claypool and Kvenvolden, 1983; Hunt, 1996). This excess water.
trend is generally attributed to the dilution of methane-rich McNeil and BeMent (1996) have hypothesized that forma-
biogenic gas with relatively methane-poor thermogenic gas tion of an almost pure methane gas occurs from residual kero-
produced at depth as temperature increases (Claypool and gen after the majority of oil has been generated and expelled. In
Kvenvolden, 1983; Whelan, 1979; Whelan and Sato, 1980). their model, alkyl groups attached to aromatic moieties in
Our experiments, however, demonstrate that at low thermal kerogen are preferentially cleaved between the first and second
maturity, the hydrocarbon fraction of thermogenic gas gener- carbon atoms leaving a methyl group behind. Reattachment of
ated directly from kerogen may also be methane-rich contain- the cleaved alkyl radical to the aromatic ring structure and
ing approximately 90 mole % methane (Fig. 6). Accordingly, subsequent cleavage of the methyl group results in generation
the amount of thermogenic methane in relatively shallow sed- of a methane-rich gas. Such a process may account for the
iments may be greater than existing models would predict. increases in the relative abundance of methane at temperatures
Although the experiments demonstrate that methane-rich gas greater than peak bitumen generation during the experiments,
is generated at low thermal maturities, it is clear that the although a dry gas is never obtained because bitumen is not
majority of volatile organically derived alteration products are expelled in our closed-system experiments. The continued pro-
formed at higher temperatures. The onset of bitumen generation duction of gas containing substantial quantities of C21 hydro-
during the experiments is accompanied by substantial increases carbons at high temperatures, however, suggests that the
in the relative amounts of C21 hydrocarbons. The composition amount of methane formed from kerogen is subordinate to the
of this gas is consistent with many natural gases associated with total hydrocarbon gas formed during the decomposition of
oil suggesting similar maturation processes in natural systems bitumen in a closed system.
(Price, 1995; Hunt, 1996; Price and Schoell, 1995). Increases in
the relative abundance of methane at high thermal maturities 4.1. Kinetic Modeling
during the experiments are also consistent with compositional
trends observed in natural systems (Rice et al., 1988; Hunt, Data for the generation of volatile hydrocarbons and carbon
1996). Despite these similarities, dry gas that is abundantly dioxide from hydrous pyrolysis experiments can be used to
present in many natural gas deposits (Mango et al., 1994) was constrain kinetic models that predict the extent and timing of
not produced during the experiments. Numerous other experi- hydrocarbon generation in sedimentary basins. In general, these
mental studies have demonstrated the persistence C21 hydro- models are based on reactions governed by first order reaction
carbons at temperatures in $400°C (Tomic et al., 1985; Do- kinetics and the Arrhenius equation, but vary significantly with
miné, 1991; Behar et al., 1992; Horsfield et al., 1992; Esser and respect to the number of reactions that are used to describe
Schwochau, 1991; Price and Schoell, 1995). Taken together, hydrocarbon generation (Lopatin, 1971, 1976; Waples, 1980;
these results demonstrate the stability of C21 hydrocarbons at Lewan, 1985; Braun and Burnham, 1990; Ungerer, 1990; Hunt
temperatures considerably higher than those typically associ- et al., 1991; Sweeney et al., 1995). Thermal maturation of
ated with the generation of natural gas. Thus, classical models kerogen to produce oil and gas is an inherently complex pro-
for the generation of dry-gas via the degradation of longer- cess involving numerous sequential and parallel reactions. Ac-
chain hydrocarbons by purely thermal processes in which suf- cordingly, even the most complex kinetic models represent
ficient energy is available to result in cleavage of all C-C bonds gross simplifications of naturally occurring processes. None-
(Tissot and Welte, 1984; Barker, 1990; Ungerer, 1990) may not theless, they represent effective tools for petroleum exploration
be appropriate. and predicting the geochemical evolution of sedimentary ba-
The apparent inability of a purely thermal model to account sins.
for the generation of dry gas has been noted previously and The Arrhenius equation may be written as follows:
resulted in a variety of new models to explain its origin and
composition (Mango et al., 1994; Price, 1995; Price and k 5 A o exp~ 2 E a/RT) (3)
Schoell, 1995; McNeil and BeMent, 1996). It has been sug-
where k is the reaction rate constant, Ao is the pre-exponential
gested that natural gas contains a significant amount of C21
factor, Ea is the activation energy, R is the ideal gas constant,
hydrocarbons at the site of generation, as is observed during
and T is temperature in Kelvin. This relationship allows data
laboratory experiments, but fractionation processes during ex-
collected during high temperature short-term experiments to be
pulsion and migration produce enrichment in methane to pro-
extrapolated to the relatively low temperatures and long reac-
duce dry gas. Other researchers have focused on chemical
tion times that typically exist in sedimentary basins. We con-
processes during generation. For example, Mango et al. (1994)
sidered applying a single activation energy model such as that
have proposed that under dry and damp conditions, the catalytic
employed by Lewan (1985) in this study, but examination of
activity of transition metals present in the Monterey Shale
gas generation trends for our experiments revealed that it would
effectively lowers the stability of hydrocarbons and increases
not be appropriate. The rate of gas generation for a reaction that
the specificity of decomposition reactions with respect to meth-
is first order with respect to the generation potential can be
ane. In the presence of excess water, however, they observed a
described by the relationship:
large decrease in the specificity of the catalyst with respect to
methane. Although transition metal catalysis cannot be un- dX/dt 5 k 2 kX (4)
equivocally demonstrated during our experiments, the latter
observation is consistent with the results of our Monterey Shale where k is the rate constant defined by Eqn. 3, t is time, and X
experiments that show production of substantial C21 hydrocar- is the fraction of the maximum amount of gas generated for a
Generation and formation of natural gas 1611

given rock (Lewan, 1985). Equation 4 indicates that at a given


temperature and pressure, the rate of gas generation is directly
proportional to the gas generation potential remaining in a rock.
Accordingly, for a single activation energy to be appropriate for
our experiments, decreases in the gas generation rate should
reflect decreases of identical magnitude in the gas generation
potential. During the experiments presented here gas generation
rates decreased by .50 % at a given temperature. Subsequent
increases in temperature, however, resulted in order of magni-
tude increases in the amount of gas generated indicating only
minor decreases in the gas generation potential had occurred at
the lower temperature. These data clearly indicate that gas
generation rates do not reflect the generation potential and are
not consistent with a single first order reaction model.
The trends observed during our experiments are consistent
with a parallel reaction model. Accordingly, we used the com-
puter code KINETICS (Braun and Burnham, 1990) to deter-
mine kinetic parameters for the generation of total low molec-
ular weight hydrocarbons (SC1-C4) and carbon dioxide. This
code can fit an average pre-exponential factor and the relative
percent of reaction associated with a given activation energy
for up to 25 first order parallel reactions. The mathematical
optimization algorithms used by KINETICS require a suffi- during which the potential for carbonate mineral precipitation
ciently large experimental dataset to fully constrain all kinetic was minimized by fluid-talc-quartz equilibrium. For the Eutaw
parameters. The large number of experiments conducted as a Shale, the entire carbon dioxide data set was used in the kinetic
function of time and temperature for the Monterey Shale al- analyses. Due to the sparseness and erratic nature of the Smack-
lowed a unique solution to be obtained. This was not the case, over Shale data set for carbon dioxide generation, kinetic
however, for the relatively sparse experimental datasets for the parameters were not retrieved.
Smackover and Eutaw shales. Consequently, to reduce the The retrieved kinetic parameters (Table 6) accurately predict
number of unknowns, pre-exponential factors for gas genera- generation of SC1-C4 hydrocarbons and carbon dioxide during
tion from these rocks were assumed to be identical to those the experiments (Figs. 1–5). For the Monterey shale a preex-
calculated for the Monterey Shale. It is important to realize, ponential factor of 1.12 3 1016s21 and a distribution of acti-
however, that the calculated activation energies are highly vation energies from 56 to 66 kcal/mole accounts for the
dependent upon this assumption. For example, a one log unit majority of low molecular weight hydrocarbon generation, with
decrease in the assumed pre-exponential factor produces a 3 the maximum percent of reaction accounted for by a value of
kcal/mole decrease in the calculated activation energy values. 66 kcal/mole. These values are within the range expected for
For the present analyses, weighting factors for the input data degradation reactions likely to produce natural gas (Ungerer,
were normalized to account for differences in the number of 1990). Using a pre-exponential factor of 1.12 3 1016s21, a
data points at a given temperature. In addition, to ensure that similar activation energy distribution was retrieved for the
relative uncertainties in the data were equally accounted for Eutaw Shale while most of the SC1-C4 hydrocarbons generated
during the regression analyses the weighting factors were from the Smackover Shale could be accounted for by an acti-
scaled to be inversely proportional the absolute amount of gas vation energy distribution ranging from 56 to 62 kcal/mole.
generated during a given experiment. Retrieval of kinetic pa- Because methane and other low molecular weight hydrocar-
rameters for each shale was performed on an absolute scale bons were still being generated at the termination of these
after normalizing the amount of gas generated at a given experiments, reactions characterized by even higher activation
temperature to the maximum generated at the highest temper- energies may have contributed significant hydrocarbon gases
ature. A 2 kcal/mole spacing between consecutive values of had temperature been increased or the experiments conducted
activation energies was used for all calculations. for longer duration. Thus, the slightly higher calculated activa-
Due to the potential for carbonate mineral precipitation dur- tion energies for SC1-C4 generation from the Monterey and
ing the experiments, data used for the kinetic analyses of Eutaw shales relative to the Smackover Shale may reflect the
organically derived carbon dioxide were carefully selected. For fact that the former experiments attained higher thermal matu-
the Monterey shale, data from experiments MS1 and MS2 at rity in terms of time and/or temperature.
225 and 275°C, respectively, were excluded from the kinetic In contrast to low molecular weight hydrocarbon generation,
analysis because inorganic carbonate was not removed from the carbon dioxide generation during heating of Monterey Shale is
starting rocks. Data from experiments MS5 and MS13 con- best accounted for by a substantially wider activation energy
ducted at 325°C were also excluded because of anomalously distribution that varies from approximately 34 to 60 kcal/mole
low carbon dioxide abundances, a likely result of carbonate and a pre-exponential factor of 3.2 3 1013s21. The calculated
mineral precipitation. Despite the exclusion of these data, car- activation energy distribution for carbon dioxide generation
bon dioxide generation at 225, 275, and 325°C during heating from the Eutaw Shale assuming a pre-exponential factor iden-
of the Monterey shale is well constrained by experiment MS15, tical to that of the Monterey Shale is characterized by a simi-
1612 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

larly wide range that varies from 36 to 56 kcal/mole. The broad Predicted trends for the generation of organically derived
activation energy distribution suggests a more diverse set of carbon dioxide at geologic heating rates are substantially dif-
reactions are responsible for the generation of carbon dioxide ferent than those indicated for the hydrocarbon fraction of
relative to SC1-C4 hydrocarbons. For example, it is likely that natural gas. Carbon dioxide generation is predicted to occur
carbon dioxide production associated with low activation en- continuously from low temperatures typical of early diagenesis
ergy values is due to the cleavage of O-bearing functional to temperatures as high as 220°C (Fig. 10). Thus, organically
groups directly from the kerogen macromolecule while that derived carbon dioxide production is not restricted to diage-
associated with high activation energies may reflect reaction of netic processes that proceed peak hydrocarbon generation as is
organic compounds with water. commonly assumed but can be expected to occur during and
Comparison of the kinetic parameters retrieved here with beyond the oil window as oil degrades to natural gas deep
previous studies shows both agreement and disagreement. within a basin.
Knauss et al. (1992) derived individual Gaussian distributions Although there are several laboratory studies reporting
of activation energies for the generation of methane, ethane, carbon dioxide generation during hydrous pyrolysis (Cooles
propane, and butane from hydrous pyrolysis experiments ex- et al., 1987; Barth et al., 1989; Burnham et al., 1992; Barth
amining the New Albany Shale. Their results indicate distribu- and Nielsen, 1993; Andresen et al., 1994, 1995; Lewan,
tions centered on a value 55 kcal/mole for the generation of all 1997), there are relatively few kinetic models that allow
gases. The lower activation energies may be due to their as- extrapolation to natural conditions. Barth and Nelson (1993)
sumption of a lower pre-exponential factor of 1 3 1014 and the report kinetic data retrieved from hydrous pyrolysis of Kim-
fact that the maximum temperature during their experiments meridge source rocks that were not treated to remove inor-
was only 330°C, substantially lower than the experiments pre- ganic carbonate. Similar experiments by Andresen et al.
sented here. Pre-exponential factors and central activation en- (1994) utilizing the same rocks, but with inorganic carbonate
ergies estimated by Barth et al. (1989) from hydrous pyrolysis removed, reveals considerably lower carbon dioxide yields
data of Kimmeridge source rocks indicate slightly lower values suggesting carbonate dissolution may have occurred to a
than those obtained here and may also be the result of lower significant extent during the experiments of Barth and Nel-
maximum thermal maturities. son (1993). Accordingly, their kinetic model likely predicts
There is a striking similarity of our results with those of carbon dioxide generation from both organic and inorganic
Horsfield et al. (1992), who used confined anhydrous pyrolysis sources and direct comparison with our kinetic model for
experiments to retrieve kinetic parameters for the generation of organically derived carbon dioxide is not appropriate.
SC1-C4 hydrocarbons during the degradation of a relatively Burnham et al. (1992) estimated activation energies ranging
immature crude oil from the North Sea. Their data yield a from 44 to 48 kcal/mole for organically derived carbon dioxide
pre-exponential factor of 1.1 3 1016 s21 and an activation generation during hydrous pyrolysis of a series of source rocks
energy distribution from 66 to 70 kcal/mole with the maximum from the Maracaibo Basin using an assumed pre-exponential
percent of reaction accounted for by a value of 66 kcal/mole. factor of 3 3 1013. Their model predicts generation of organ-
The consistency of these results suggests that similar reactions ically derived carbon dioxide commences at 80°C and ceases at
are responsible for the generation of low molecular weight 140°C, considerably lower than our kinetic data predicts (Fig.
hydrocarbons in both sets of experiments. That the rate of 10). However, because inorganic carbonate was not removed
SC1-C4 generation in our experiments is characterized by sim- from their initial source rocks, Burnham et al. (1992) were
ilar rates despite occurring via a two-step process involving the forced to assume that carbon dioxide generated at high tem-
generation of bitumen from kerogen and its subsequent degra- perature during their experiments was inorganic in origin while
dation to form low molecular weight hydrocarbons suggests the that generated at low temperature was organic in origin. Based
degradation of bitumen is the rate-limiting step that dominates on this assumption, high-temperature carbon dioxide was ex-
the timing and extent of gas generation. cluded from their kinetic analysis. Generation of carbon diox-
The relative timing of oil and gas generation are more readily ide at high thermal maturities during the experiments presented
apparent when the kinetic parameters retrieved from experi- here suggests that such an assumption may not be valid. More-
mental data are applied to temperature and time conditions over, it is likely that carbonate mineral dissolution and/or
typical of petroleum producing sedimentary basins. Gas gen- precipitation was occurring at all temperatures during their
eration curves for a heating rate of 8°C/My indicate that the experiments precluding unequivocal determination of the
bulk of SC1-C4 production is restricted to 160 to 230°C for the amounts of organically derived carbon dioxide as a function of
Monterey and Eutaw shales while generation from the Smack- time and temperature.
over Shale ceases at approximately 200°C (Fig. 10). These Extrapolation of experimental results to constrain natural gas
temperatures are considerably higher than the 60 to 160°C generation in nature suggests both hydrocarbons and carbon
temperature range typically invoked for the oil window (Tissot dioxide are likely to be generated during the degradation of oil
and Welte, 1984; Hunt, 1996). As has been previously noted and kerogen deep within a sedimentary basin. It is important to
(Sweeney et al., 1987; Horsfield et al., 1992), the chemical remember that the maximum predicted temperatures for gas
reactions responsible for oil generation may be temporally and generation (approximately 230°C) are directly constrained by
spatially isolated from those responsible for degradation and the time and temperature conditions of the experiments used to
are not characterized by a continuous transition. This is espe- constrain the kinetic models. Significant carbon dioxide and
cially true in case of the S-rich Monterey shale for which oil methane production were still occurring at the termination of
generation is predicted to be complete at 100°C for a heating the experiments indicating that the maximum gas potential of
rate of 8°C/My (Hunt et al., 1991). the organic matter had not been exhausted. Thus, it is likely that
Generation and formation of natural gas 1613

Fig. 10. Predicted gas generation curves for (a) SC1-C4 hydrocarbons and (b) carbon dioxide from the Monterey,
Smackover, and Eutaw shales in a hypothetical basin characterized by a heating rate of 8°C/My. The kinetic parameters in
Table 6 and the computer code KINETICS (Braun and Burnham, 1990) were used for the construction of this diagram. Also
shown for comparison is the generation curve for carbon dioxide predicted using the kinetic parameters of Burnham et al.
(1992) applied to the generation potential determined during the Monterey Shale experiments.

in nature, carbon dioxide and methane production may continue Land, 1986; Jenden et al., 1988a,b; Smith and Ehrenberg, 1989;
deep within a basin to temperatures in excess of 230°C. Poulson et al., 1995), considerably lower than the minimum
value of 47 mole % observed during hydrous pyrolysis of the
4.2. Controls on the Abundance of Carbon Dioxide in Monterey, Eutaw, and Smackover shales (Fig. 6). The relative
Natural Systems abundance of carbon dioxide decreased systematically with
increasing thermal maturity during the experiments and it is
The abundance of carbon dioxide in natural gas shows large likely that lower mole % carbon dioxide values may have been
and systematic variations with values ranging from 0 to 100 observed had the experiments been of longer duration. How-
mole %. In many cases, carbon dioxide abundances are char- ever, even if all sedimentary organic matter remaining at the
acterized by a strong positive correlation with increasing tem- end of each experiment were to form methane with continued
perature (Smith and Ehrenberg, 1989; Franks and Forester, reaction, the large amounts of carbon dioxide generated earlier
1984; Lundegard and Land; 1986; James, 1990). In general, the preclude attainment of relative carbon dioxide abundances as
majority of reported carbon dioxide abundances in natural gas low as those observed in the majority of natural gas samples.
are ,10 mole % (Franks and Forester, 1984; Lundegard and Thus, comparison of experimental data with natural gas com-
1614 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

positions suggests that carbon dioxide is being removed from


thermogenic gas in natural systems by secondary processes.
Sinks for carbon dioxide in sedimentary basins include pref-
erential dissolution in aqueous pore fluids and subsequent mi-
gration, and/or precipitation as carbonate minerals. Although it
is likely that the former may occur to a significant extent, it
cannot adequately account for the strong correlation of relative
carbon dioxide abundance with subsurface temperature. Con-
vincing arguments have been made to suggest that this trend
results from the control of carbon dioxide partial pressures
(PCO2) by thermodynamic equilibrium involving feldspar, clay,
and carbonate minerals (Smith and Ehrenberg, 1989; Hutcheon
and Abercrombie, 1990; Hutcheon et al., 1993). Smith and
Ehrenberg (1989) suggest that at shallow levels within a sedi-
mentary basin where temperatures are ,120°C, precipitation of
carbonate minerals will occur while at deeper levels, where
temperatures exceed 120°C, dissolution of sedimentary carbon-
ate will occur. This model is based on the assumption that
carbon dioxide generation is an early stage phenomenon that
Fig. 11. Predicted alteration mineral assemblage as a function of rock
produces excess carbon dioxide at shallow levels while at total organic carbon content during reaction of a seawater salinity fluid
deeper levels the carbon dioxide potential of kerogen is ex- with a hypothetical rock initially containing 30 wt% plagioclase (mole
hausted and dissolution of sedimentary carbonate occurs to fraction anorthite 5 0.3), 10 wt% K-feldspar, and 60 wt% quartz at
maintain saturation with respect to calcite. Continuous carbon 200°C and 350 bars. See text for model details.
dioxide generation before, during, and after peak bitumen gen-
eration during the experiments presented here, however, sug-
gests that carbon dioxide partial pressures may remain above amount of carbon dioxide generated during hydrous pyrolysis
those required to maintain saturation with calcite even at rela- of Monterey shale (8.0 mmole CO2/g TOC) was used to di-
tively high levels of thermal maturity, provided there is suffi- rectly relate the carbon dioxide quantities to sedimentary TOC
cient TOC content in the sediments. Evidence in support of this contents.
concept is provided by relatively depleted d13C values for Variations in the predicted mineralogy as a function of
carbon dioxide gas and carbonate cements in some basins that increasing TOC are shown in Fig. 11. In the absence of carbon
indicate an organic origin at both shallow and deep levels dioxide (0 wt% TOC), anorthite is thermodynamically unstable
(Franks and Forester, 1984; Smith and Ehrenberg, 1989). In the and prehnite is predicted as the only calcic phase. Upon intro-
Texas Gulf Coast, Franks and Forester (1984) report d13C duction of TOC and its associated carbon dioxide to the system,
values for carbonate cements that decrease systematically with calcite formation commences almost immediately. Owing to
depth as temperature increases to values as high as 210°C, the low solubility of calcite, PCO2 values in equilibrium with the
consistent with increased generation of organically derived stable alteration assemblage consisting of K-feldspar, albite,
carbon dioxide at depth. Thus, field data provide evidence for muscovite, prehnite, calcite, quartz assemblage are buffered at
carbon dioxide generation and removal from natural gas as 0.25 bars (Fig. 12). Continued increases in TOC result in the
calcite at shallow and deep levels within a sedimentary basin. destruction of prehnite as Ca is cycled into calcite. Low PCO2
A requirement of the above hypothesis, however, is that suffi- values are maintained until prehnite is exhausted as a reactant,
cient Ca from noncarbonate sources such as aluminosilicate at which point calcite formation ceases and carbon dioxide
minerals exists within the rock to form calcite. accumulates in solution. Because the amount of prehnite
To test the above hypothesis we conducted a series of the- present is directly limited by the availability of Ca present in
oretical reaction path calculations using the EQ3NR/EQ6 com- the initial plagioclase, buffering of PCO2 due to calcite precip-
puter code (Wolery, 1992; Wolery and Daveler, 1992) and a itation continues to higher TOC contents for rocks containing
thermodynamic database generated at 350 bars using greater amounts of plagioclase. These calculations suggest that
SUPCRT92 (Johnson et al., 1992). The goal of these calcula- in many sedimentary basins associated with the production of
tions was to evaluate within a thermodynamic framework the oil and natural gas, typical plagioclase abundances are suffi-
amounts of noncarbonate Ca required to effectively remove cient to facilitate calcite formation and PCO2 buffering in all but
carbon dioxide as calcite from natural gas. The calculations the most organic-rich source rocks. For example, a rock con-
involved reacting a hypothetical rock containing 10 wt% K- taining 30 wt% plagioclase would require a TOC content .4
feldspar and varying amounts of plagioclase and quartz, with a wt% to exceed the PCO2 buffer capacity (Fig. 12). Moreover,
Na-Ca-Cl fluid of seawater salinity at 200°C and 350 bars. A this TOC content represents a minimum value since in natural
plagioclase composition of 0.3 mole fraction anorthite systems kerogen may not have attained thermal maturation
(CaAl2Si2O8) was used for these calculations. An initial fluid/ levels equivalent to those reached during the 375°C phase of
rock mass ratio of 0.32 was chosen which corresponds to an the Monterey Shale experiment which was used as the input
initial porosity of approximately 10 %. Because kerogen cannot constraint for carbon dioxide potential during these calcula-
be treated within a thermodynamic framework, CO2(g) was tions.
added to the system as an additional reactant. The maximum Considering the abundance of Ca in sedimentary rocks, it is
Generation and formation of natural gas 1615

responsible for its production. Carbon dioxide is the dominant


volatile organic alteration product during the heating of rocks
under hydrous conditions and represents an important alteration
product during the decomposition of bitumen and kerogen in
the presence of water. Requisite oxygen for carbon dioxide
formation is likely derived from kerogen at low maturities and
the disproportionation of water at high maturities. Hydrogen
released during the disproportionation of water is likely con-
sumed by hydrocarbon formation.
Kinetic analyses of the experimental results suggest gener-
ation of SC1-C4 hydrocarbons from the Monterey, Smackover,
and Eutaw shales can be accounted for by a narrow range of
activation energies while carbon dioxide production is charac-
terized by a substantially broader distribution. Application of
retrieved kinetics parameters to times and temperatures typical
of subsiding sedimentary basins indicates that generation of
volatile hydrocarbons is restricted to relatively high thermal
maturities while carbon dioxide generation occurs at both low
Fig. 12. Predicted PCO2 in equilibrium with a hypothetical rock and high thermal maturities. Thus, in contrast to the bulk of
containing varying amounts of plagioclase (solid isopleths), 10 wt% C1-C4 generation which is predicted to occur following peak
K-feldspar, and 60 wt% quartz at 200°C and 350 bars as a function of bitumen generation, production of carbon dioxide is predicted
rock total organic carbon content. A plagioclase composition of 0.3
to occur before, during, and after the generation of liquid
mole fraction anorthite was used for the construction of this diagram.
See text for model details. hydrocarbons.
Theoretical reaction path modeling can be used to place
constraints on sinks for carbon dioxide in natural systems.
Calculations suggest that precipitation of calcite is an effective
likely that calcite precipitation will influence the composition
means to remove carbon dioxide from natural gas provided Ca
of most natural gas samples resulting in carbon dioxide abun-
is available in the rock. Because generation of carbon dioxide
dances that are substantially lower than those observed during
may continue to high thermal maturities, precipitation of or-
hydrous pyrolysis experiments. Widespread carbonate precip-
ganically derived carbonate cements is likely at deep levels
itation in response to thermogenic carbon dioxide production,
within sedimentary basins. Thus, the experimental results in
however, will not necessarily result in the production of calcite-
conjunction with theoretical modeling suggest that high relative
rich rocks that are easily identifiable. Complete precipitation of
carbon dioxide abundances are not anomalous but may be
thermogenic carbon dioxide in a rock containing 4 wt% TOC
characteristic of pristine natural gas, while low carbon dioxide
will produce only 3.31 wt% calcite which corresponds to 0.39
abundance may be indicative of secondary reactions that results
wt% carbonate-C. This relatively low carbonate abundance
in the precipitation of carbonate minerals.
would be indistinguishable from that in numerous marine sed-
imentary rocks that have not experienced thermogenic carbon
dioxide production. Acknowledgments—The authors would like to thank Dr. Michael Le-
Results of hydrous pyrolysis experiments and model calcu- wan for providing the Monterey Shale sample and the Louisiana Basin
Research Group for the Smackover and Eutaw shale samples. Organic
lations indicate that the formation of CO2-rich natural gas does petrographic analyses by Lorraine Eglinton are gratefully appreciated.
not require the addition of carbon dioxide from an external Thoughtful reviews by Alan Burnham, Ian Hutcheon, and an anony-
source such as igneous activity or by other processes such as mous reviewer substantially improved this manuscript. Funds for this
thermochemical sulfate reduction (Krouse et al., 1988). In project were provided by U.S. Department of Energy grant numbers
rocks characterized by high TOC and/or low plagioclase abun- DE-FGO2-86ER13466, DE-FGO2-92ER14232, and DE-FGO2-
97ER14746 and by Gas Research Institute contract numbers 5088-260-
dance the buffering capacity of the rock will be exceeded, and 1746 and 5091-260-2298.
PCO2 will be characterized by relatively high values. Thus, the
results of laboratory experiments in conjunction with reaction Editorial handling: J. T. Senftle
path modeling suggests that organically derived pristine natural Accepted for publication after revision: G. Faure
gas is CO2-rich and that secondary reactions involving inor-
ganic sedimentary rock components may be required to pro-
REFERENCES
duce methane-rich gas observed in most deposits.
Acholla F. V. and Orr W. L. (1993) Pyrite removal from kerogen
5. SUMMARY without altering organic matter: the chromous chloride method.
Energy & Fuels 7, 406 – 410.
Laboratory experiments represent an effective means to con- Andresen B., Throndsen T., Barth T., and Bolstad J. (1994) Thermal
strain processes responsible for organic transformations that generation of carbon dioxide and organic acids from different source
occur in sedimentary basins. Results of hydrous pyrolysis ex- rocks. Org. Geochem. 21, 1229 –1242.
Andresen B., Throndsen T., Råheim A., and Bolstad J. (1995) A
periments involving Monterey, Smackover, and Eutaw shales comparison of pyrolysis products with models for natural gas gen-
indicate that large and systematic variations in the composition eration. Chem. Geol. 126, 261–280.
of natural gas occur in response to changes in the reactions Barker C. (1990) Calculated volume and pressure changes during the
1616 J. S. Seewald, B. C. Benitez-Nelson, and J. K. Whelan

thermal cracking of oil to gas in reservoirs. AAPG Bull. 74, 1254 – Hoering T. C. (1984) Thermal reaction of kerogen with added water,
1261. heavy water, and pure organic substances. Org. Geochem. 5, 267–
Barth T. and Nielsen S. B. (1993) Estimating kinetic parameters for 278.
generation of petroleum and single compounds from hydrous pyrol- Horsfield B., Schenk H. J., Mills N., and Welte D. H. (1992) Closed-
ysis of source rocks. Energy & Fuels 7, 100 –110. system programmed-temperature pyrolysis for simulating the con-
Barth T., Borgund A. E., and Hopland A. L. (1989) Generation of version of oil to gas in a deep petroleum reservoir: Compositional
organic compounds by hydrous pyrolysis of Kimmeridge oil shale - and kinetic findings. Org. Geochem. 19, 191–204.
bulk results and activation energy calculations. Org. Geochem. 14, Hunt J. M. (1996) Petroleum Goechemistry and Geology. W.H. Free-
69 –76. man.
Behar F., Kressman J. L., Rudkiewicz J. L., and Vandenbroucke M. Hunt J. M., Lewan M. D., and Hennet R. J.-C. (1991) Modeling oil
(1992) Experimental simulation in a confined system and kinetic generation with time-temperature index graphs based on the Arrhe-
modeling of kerogen and oil cracking. Org. Geochem. 19, 173–189. nius equation. AAPG Bull. 75, 795– 807.
Bischoff J. L. and Rosenbauer R. J. (1985) An empirical equation of Hutcheon I. and Abercrombie H. (1990) Carbon dioxide in clastic rocks
state for hydrothermal seawater (3.2 percent NaC1). Amer. J. Sci. and silicate hydrolysis. Geology 18, 541–544.
285, 725–763. Hutcheon I., Shevalier M., and Abercrombie H. J. (1993) pH buffering
Bonham L. C. (1978) Solubility of methane in water at elevated by metastable mineral-fluid equilibria and evolution of carbon diox-
temperatures and pressures. AAPG Bull. 49, 2478 –2481. ide fugacity during burial diagenesis. Geochim. Cosmochim. Acta
Bowers T. S. and Helgeson H. C. (1983) Calculation of the thermody- 57, 1017–1027.
namic and geochemical consequences of nonideal mixing in the James A. T. (1990) Correlation of reservoired gases using the carbon
system H2O-CO2-NaCl on phase relations in geologic systems: isotopic compositions of wet gas components. AAPG Bull. 74, 1441–
Equation of state for H2O-CO2-NaCl fluids at high pressures and 1458.
temperatures. Geochim. Cosmochim. Acta 47, 1247–1275. Jenden P. D., Kaplan I. R., Poreda R. J., and Craig H. (1988a) Origin
Braun R. L. and Burnham A. K. (1990) KINETICS: A computer of nitrogen-rich natural gases in the California Great Valley; evi-
program to analyze chemical reaction data. Lawrence Livermore dence from helium, carbon and nitrogen isotope ratios. Geochim.
National Laboratory Report UCID-21588, Rev. 1. Cosmochim. Acta 52, 851– 861.
Burnham A. K., Braun R. L., Gregg H. R., and Samoun A. M. (1987) Jenden P. D., Newell K. D., Kaplan I. R., and Watney W. L. (1988b)
Comparison of measuring kerogen pyrolysis rates and fitting kinetic Composition and stable-isotope geochemistry of natural gases from
parameters. Energy & Fuels 1, 452– 458. Kansas, midcontinent, U.S.A. Chem. Geol. 71, 117–147.
Burnham A. K., Braun R. L., Sweeney J. J., Reynolds J. G., Vallejos Johnson J. W., Oelkers E. H., and Helgeson H. C. (1992) SUPCRT92:
C., and Talukdar S. (1992) Kinetic modeling of petroleum formation Software package for calculating the standard molal thermodynamic
in the Maracaibo basin: final report. U.S. Department of Energy properties of minerals, gases, aqueous species, and reactions among
Report DOE/BC/92001051, Bartlesville, Oklahoma. them as functions of temperature and pressure. Computers Geosci.
Canfield D. E., Raiswell R., Westrich J. T., Reaves C. M., and Berner 18, 899 –947.
R. A. (1986) The use of chromium reduction in the analysis of Karlsen D. A. and Larter S. R. (1991) Analysis of petroleum fractions
reduced inorganic sulfur in sediments and shales. Chem. Geol. 54, by TLC-FID: applications to petroleum reservoir description. Org.
149 –155. Geochem. 17, 603– 617.
Claypool G. E. and Kvenvolden K. A. (1983) Methane and other Karweil J. (1955) Die Metamorphose der kohlen vom standpunkt der
hydrocarbon gases in marine sediment. Ann. Rev. Earth Planet. Sci. physikalischen chemie. Z. Deutschen Geolog. Gesell. 107, 132–139.
11, 299 –3217. Knauss K. G., Copenhaver S. A., Braun R. L., and Burnham A. K.
Cooles G. P., Mackenzie A. S., and Quigley T. M. (1986) Calculation (1992) Hydrous pyrolysis of New Albany and Phosphoria Shales:
of petroleum masses generated and expelled from source rocks. Org. Effects of temperature and pressure on the kinetics of production of
Geochem. 10, 235–245. carboxylic acids and light hydrocarbons. Amer. Chem. Soc. Div. of
Cooles G. P., Mackenzie A. S., and Parkes R. J. (1987) Non-hydro- Fuel Chem. 37, 1621–1625.
carbons of significance in petroleum exploration: volatile fatty acids Krouse H. R., Viau C. A., Eliuk L. S., Ueda A., and Halas S. (1988)
and non-hydrocarbon gases. Mineral. Mag. 51, 483– 493. Chemical and isotopic evidence of thermochemical sulphate reduc-
Dominé F. (1991) High pressure pyrolysis of n-hexane, 2,4-dimeth- tion by light hydrocarbon gases in deep carbonate reservoirs. Nature
ylpentane and 1-phenylbutane: Is pressure an important geochemical 333, 415– 419.
parameter? Org Geochem. 17, 619 – 634. Leif R. N. and Simoneit B. R. T. (1995) Ketones in hydrothermal
Eglinton T. I. and Douglas A. G. (1988) Quantitative study of biomar- petroleums and sediment extracts from Guaymas Basin, Gulf of
ker hydrocarbons released from kerogens during hydrous pyrolysis. California. Org. Geochem. 23, 889 –904.
Energy & Fuels 2, 81– 88. Lewan M. D. (1985) Evaluation of petroleum generation by hydrous
Espitalié J., Laporte J. L., Madec M., Marquis F., Leplat P., and Paulet pyrolysis experimentation. Phil. Trans. Roy. Soc. London A315,
J. (1977) Méthode rapide de caractérisation des roches mères, de leur 123–134.
potential pétrolier et de leur degré d’évolution. Inst. Fr. Pétr. Rev. Lewan M. D. (1993a) Laboratory simulation of petroleum formation:
32, 23– 43. Hydrous pyrolysis. In Organic Geochemistry (ed. M.H. Engel and
Espitalié J., Marquis F., and Barsony I. (1984) Geochemical logging. In S. A. Macko), pp. 419 – 442. Plenum.
Analytical Pyrolysis (ed. K. J. Voorhees), pp. 276 –304. Butter- Lewan M. D. (1993b) Primary oil migration and expulsion as deter-
worths. mined by hydrous pyrolysis. In Proceedings 13th World Petroleum
Esser W. and Schwochau K. (1991) Evolution of individual hydrocar- Congress, Buenos Aires, 1991, pp. 215–223. Wiley.
bons (C1-C4) by nonisothermal pyrolysis of petroleum source rocks. Lewan M. D. (1997) Experiments on the role of water in petroleum
J. Analyt. Appl. Pyrolysis 22, 61–71. formation. Geochim. Cosmochim. Acta 61, 3691–3723.
Franks S. G. and Forester R. W. (1984) Relationships among secondary Lopatin N. V. (1971) Temperature and geologic time as factors of
porosity, pore-fluid chemistry and carbon dioxide, Texas Gulf Coast. cabonification. Izvest. Akad. Nauk SSSR, Seriya Geologich. 3, 95–
In Clastic Diagenesis (ed. D.A. McDonald and R.C. Surdam); AAPG 106.
Memoir 37, pp. 63– 80. AAPG. Lopatin N. V. (1976) The determination of the influence of temperature
Haar L., Gallagher J., and Kell G. (1980) Thermodynamic properties of and geologic time on the catagenic processes of coalification and
fluid water. In Proceeding of the 9th International Conference on the oil-gas formation. In Issledovaniya organicheskogo veshchestva
Properties of Steam, Munich 1979 (ed. J. Straub and K. Scheffler), sovremennykh i iskopaemykh osadkov (Research on organic matter
pp. 69 – 82. Pergamon. of modern and fossil deposits), pp. 361–366. Moscow, Akademii
Helgeson H. C., Knox A. M., Owens C. E., and Shock E. L. (1993) Nauk SSSR, Isdatek’stvo, Nauka.
Petroleum, oil field waters, and authigenic mineral assemblages: Are Lundegard P. D. and Land L. S. (1986) Carbon dioxide and organic
they in metastable equilibrium in hydrocarbon reservoirs? Geochim. acids: their role in porosity enhancement and cementation, Paleogene
Cosmochim. Acta 57, 3295–3339. of the Texas Gulf Coast. In Roles of Organic Matter in Sediment
Generation and formation of natural gas 1617

Diagenesis (ed. D.L. Gautier); Soc. Econ. Pal. Mineral. Spec. Publ. Seyfried W. E., Jr., Janecky D. R., and Berndt M. E. (1987) Rocking
38, 129 –146. autoclaves for hydrothermal experiments II. The flexible reaction-
Mango F. D., Hightower J. W., and James A. T. (1994) Catalysis in the cell system. In Experimental Hydrothermal Techniques (ed. G. C.
origin of natural gas. Nature 368, 536 –538. Ulmer and H. L. Barnes), pp. 216 –240. Wiley.
McNeil R. I. and BeMent W. O. (1996) Thermal stability of hydrocar- Siskin M. and Katritzky A. R. (1991) Reactivity of organic compounds
bons: laboratory criteria and field examples. Energy & Fuels 10, in hot water: Geochemical and technological implications. Science
60 – 67. 254, 231–237.
Monthioux M. (1988) Expected mechanisms in nature and in confined- Smith J. T. and Ehrenberg S. N. (1989) Correlation of carbon dioxide
system pyrolysis. Fuel 67, 843– 847. abundance with temperature in clastic hydrocarbon reservoirs: rela-
Monthioux M., Landais P., and Monin J.-C. (1985) Comparison be- tionship to inorganic chemical equilibrium. Mar. Petrol. Geol. 6,
tween natural and artificial maturation series of humic coals from the 129 –135.
Mahakam Delta, Indonesia. Org. Geochem. 8, 275. Stalker L., Farrimond P., and Larter S. R. (1994) Water as an oxygen
Nelson B. C., Eglinton T. I., Seewald J. S., Vairavamurthy A., and source for the production of oxygenated compounds (including CO2
Miknis F. P. (1995) Transformations in organic sulfur speciation precursors) during kerogen maturation. Org. Geochem. 22, 477– 486.
during maturation of Monterey shale: Constraints from laboratory Sweeney J. J., Burnham A. K., and Braun R. L. (1987) A model of
experiments. In Geochemical Transformations of Sedimentary Sulfur hydrocarbon generation from type I kerogen: Application to the
(ed. M.A. Vairavamurthy and M.A.A. Schoonen); ACS Symposium Vinta Basin, Utah. Bull. AAPG 71, 967–985.
Series 612, pp. 138 –166. Amer. Chem. Soc. Sweeney J. J., Talukdar S., Braun R. L., Burnham A., and Vallejos C.
Poulson S. R., Ohmoto H., and Ross T. P. (1995) Stable isotope (1995) Chemical kinetic model of hydrocarbon generation, expul-
geochemistry of waters and gases (CO2, CH4) from the overpres- sion, and destruction applied to the Maracaibo Basin, Venezuela..
sured Morganza and Moore-Sams fields, Louisiana Gulf Coast. Appl. AAPG Bull. 79, 1515–1532.
Geochem. 10, 407– 417. Tannenbaum E. and Kaplan I. R. (1985) Low-MT hydrocarbons gen-
Price L. C. (1981) Aqueous solubility of crude oil to 400C̊ and 2,000
erated during hydrous and dry pyrolysis of kerogen. Nature 317,
bars pressure in the presence of gas. J. Petrol. Geol. 4, 195–223.
708 –709.
Price L. C. (1993) Thermal stability of hydrocarbons in nature: Limits,
Tissot B. P. and Welte D. H. (1984) Petroleum Formation and Occur-
evidence, characteristics, and possible controls. Geochim. Cosmo-
rence. Springer-Verlag.
chim. Acta 57, 3261–3280.
Tomic J., Behar F., Vandenbroucke M., and Tang Y. (1995) Artificial
Price L. C. (1995) Origins, characteristics, controls, and economic
maturation of Monterey kerogen (Type II-S) in a closed system and
viabilities of deep-basin gas resources. Chem. Geol. 126, 335–349.
Price L. C. and Schoell M. (1995) Constraints on the origins of comparison with Type II kerogen: implications on the fate of sulfur.
hydrocarbon gas from compositions of gases at their site of origin. Org. Geochem. 23, 647– 660.
Nature 378, 368 –371. Ungerer P. (1990) State of the art research in kinetic modelling of oil
Rice D. D., Threlkeld C. N., and Vuletich A. K. (1988) Analyses of formation and expulsion. Org. Geochem. 16, 1–25.
natural gases from Anadarko Basin, southwestern Kansas, western Waples D. W. (1980) Time and temperature in petroleum formation:
Oklahoma, and Texas Panhandle. USGS Open File Rep. 88-391. application of Lopatin’s method to petroleum exploration. AAPG
Saccocia P. J. and Seyfried W. E., Jr. (1990) Talc-quartz equilibria and Bull. 64, 916 –926.
the stability of magnesium chloride complexes in NaCl-MgCl2 so- Whelan J. K. (1979) C1 to C7 hydrocarbons from IPOD holes 397 and
lutions at 300, 350, and 400°C, 500 bars. Geochim. Cosmochim. Acta 397A. In Initial Reports of the Deep Sea Drilling Project (ed. U. Rad
54, 3283–3294. et al.), Vol. 47(1), 531–539. U.S. Govt. Printing Office.
Schenk H. J. and Horsfield B. (1993) Kinetics of petroleum generation Whelan J. K. and Sato S. (1980) C1-C5 hydrocarbons from core gas
by programmed-temperature closed - versus open-system pyrolysis. pockets, Deep Sea Drilling Project Legs 56 and 57, Japan Trench
Geochim. Cosmochim. Acta 57, 623– 630. Transect. In Initial Reports of the Deep Sea Drilling Project (ed.
Seewald J. S. (1994) Evidence for metastable equilibrium between E. Horan et al.), Vol. 56-57(2), 1335–1347. U. S. Govt. Printing
hydrocarbons under hydrothermal conditions. Nature 370, 285–287. Office.
Seewald J. S. (1996) Mineral redox buffers and the stability of organic Wolery T. J. (1992) EQ3NR, A Computer Program for Aqueous Spe-
compounds under hydrothermal conditions. Mat. Res. Soc. Symp. ciation-Solubility Calculations: Theoretical Manual, User’s Guide,
Proc. 432, 317–331. and Related Documentation. Lawrence Livermore National Lab.
Seewald J. S. and Seyfried W. E., Jr. (1990). The effect of temperature Wolery T. J. and Daveler S. A. (1992) EQ6, A Computer Program for
on heavy and base metal mobility in subseafloor hydrothermal sys- Reaction Path Modelling of Aqueous Geochemical Systems: Theo-
tems: constraints from basalt alteration experiments and field studies. retical Manual, User’s Guide, and Related Documents. Lawrence
Earth Planet. Sci. Lett. 101, 388 – 403. Livermore National Lab.

You might also like