Sei sulla pagina 1di 20

Organic Matter: Chapter 16

Organic Matter: Productivity, Accumulation, and Preservation in Recent and Ancient


Sediments, by Jean K. Whelan and John W. Farrington

16. Natural Hydrous Pyrolysis: Petroleum Generation in


Submarine Hydrothermal Systems

The conversion of organic matter to petroleum products by hydrothermal activity is a geologically


rapid process, occurring in nature in many types of submarine environments. Geologically immature
organic matter of marine sediments is being altered to petroleum by this process, which is
analogous to laboratory hydrous pyrolysis. This activity has been studied in Guaymas Basin (Gulf of
California), Escanaba Trough and Middle Valley (NE Pacific), Bransfield Strait (Antarctica), and
Atlantis II Deep (Red Sea). Petroleum-like products are formed by the same process from
contemporary organic detritus and/or viable microorganisms when they become entrained by
turbulent mixing into the discharging vent waters, resulting in an instantaneous hydrous pyrolysis.
This latter case is ubiquitous in hydrothermal systems and has been studied in sediment-starved
fields of hydrothermal vents emanating directly from oceanic ridges, such as on the East Pacific
Rise at 13°N and 21°N and on the Mid-Atlantic Ridge at 26°N. The hydrocarbon products (methane
to asphalt) generated in all these areas have been compared in terms of composition, organic
matter sources, and analogy to reservoir petroleum. Preliminary data from laboratory hydrous
pyrolysis studies indicate that some of the organic matter interconversions observed in nature can
be duplicated and thereby studied in greater detail. Hydrous pyrolysis is a natural process in
hydrothermal vent systems generating petroleum, and the associated fluids are efficient solvents for
primary and secondary migration of that petroleum. In addition, such petroleum represents a major
input of carbon to the primary chemosynthetic productivity around hydrothermal vent systems.

Organic matter of marine sedimentary basins is derived from the syngenetic residues of biogenic
debris that originates from both autochthonous (marine) and allochthonous (continental) sources (e.
g., Simoneit 1978, 1982a; Hunt 1979). The preservation of organic matter in sediments depends on
the initial diagenetic processes, which involve microbial degradation and chemical conversions,
together with the acidity and redox potential of the environment (e.g., Didyk et al. 1978; Demaison
and Moore 1980). Subsequent sediment maturation and lithification cause metamorphism of the
organic matter, ultimately yielding petroleum products through the effects of temperature, pressure,
and petrology (Hunt 1979; Tissot and Welte 1984). However, the action of hydrothermal processes
on such sedimentary organic matter has been found to generate petroleum products in Guaymas
Basin almost "instantaneously" in terms of geological time (Simoneit and Lonsdale 1982). The
present paper reviews this process using examples from the extensively studied Guaymas Basin
and other areas of hydrothermal activity at both sedimented and bare-rock spreading centers and
compares the products with those obtained from laboratory simulation using hydrous pyrolysis.

The analytical techniques of organic geochemistry have been used extensively to examine the
character of organic matter in the geologic record in terms of its structural and compositional
makeup (e.g., Simoneit 1978; van de Meent et al. 1980; Tissot and Welte 1984; Johns 1986). The
sources, diagenetic and catagenetic histories, and migration mechanisms of this organic matter can
be evaluated from such data. In the following discussion, organic matter is classified as gas, lipids

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (1 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

(bitumen), asphaltenes, and kerogen (with the kerogen precursor, sometimes called
"pseudokerogen" or "protokerogen"), and the characteristics of this organic matter have been
described elsewhere (e.g., Simoneit 1983, 1988).

Nature of Organic Matter in Maturing Basins

Immature organic matter in recent sediments is composed of minor amounts (based on total organic
carbon content) of biogenic gas (CH4 and CO2, sometimes H2S), significant lipid residues of
terrigenous and/or marine origins, and a major macromolecular fraction consisting of fulvic and
humic acids and particulate detritus (e.g., biopolymer fragments, cell membranes, and
miscellaneous carbonaceous matter). The lipids and macromolecular material undergo alteration
and diagenesis (both thermal and microbial) according to the environmental conditions during
transport and deposition, i.e., oxidative degradation in high-energy, oxygenated environments and
reductive changes in low-energy anaerobic environments (Didyk et al. 1978; Demaison and Moore
1980; Simoneit 1983).

After early diagenesis, thermal maturation of organic matter commences at about 50°C (Hunt 1979)
as burial increases with a concomitant rise in temperature due to the geothermal gradient. This
process produces some low-temperature cracking products from the kerogen (nonextractable
macromolecular organic matter), such as natural gas (C1-C8+) and bitumen (C8-C40+, including a
large envelope of an unresolved complex mixture of compounds or UCM), which are added to the
endogenous biogenic gas and lipid residues. Maturation of protokerogen is generally believed to
occur via molecular rearrangements and addition of geomonomers by copolymerization. Further
heating during catagenesis (to about 150°C) generates additional amounts of bitumen and gas,
which far exceed the original concentrations of endogenous lipids and gas, thus erasing the bulk
compositional signatures of the latter and resulting in the characteristic distributions of petroleum
compounds (Hunt 1979; Simoneit 1983; Tissot and Welte 1984). Late stages of catagenesis and
the subsequent high-temperature alteration phase called metagenesis (very deep burial) primarily
generate methane from both bitumen and kerogen.

Nature of Organic Matter in Hydrothermal Systems

Hydrothermal systems can also act on sedimentary organic matter and on entrained ambient
organic detritus in water or in talus; these processes result in "instantaneous" diagenesis and
catagenesis, thus producing petroleum products analogous to those from the slower acting
geothermal changes described above (Simoneit and Lonsdale 1982; Simoneit 1983; 1984a,b; 1985;
1988; 1990). Gas (C1-C8+, CO2 and H2S) and bitumen (C8-C40+, with a large UCM) are cracked
from the pseudokerogen and biopolymers and are added to the endogenous gas and lipids. The
bitumen additionally contains products characteristic of elevated thermal processes (e.g.,
polynuclear aromatic hydrocarbons or PAH, stabilized molecular markers such as 17 (H)-hopanes,
etc.). The spent kerogen remains as disseminated, amorphous "activated" carbon (Curray et al.
1982; Simoneit 1982b; Simoneit et al. 1984).

The effects of pressure, temperature, and time on the chemistry of the organic matter are all
interrelated. Temperature and time, which primarily effect petroleum generation, control
mineralogical interactions and the products derived from the various compositions of kerogen.
Migration processes are understood less thoroughly. In sedimentary basins, migration of petroleum
seems to occur primarily by oil phase flow; migration by molecular diffusion and in water solution is

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (2 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

minor (Hunt 1979; Tissot and Welte 1984; Durand 1988). In hydrothermal areas, migration
proceeds with extreme efficiency by thermally driven diffusion, in solution (H2O/CH4/CO2
cosolvent) and by advection and mass transport as oil and/or emulsions (Kawka and Simoneit
1987; Simoneit, Kawka, and Brault 1988; Didyk and Simoneit 1989; Simoneit et al. 1990). Despite
these differences, the overall result may be the same for both regimes. Pressure in both cases aids
mainly the solubility of the petroleum in the H2O/CH4/CO2 fluids (Price, Clayton, and Rumen 1981;
Price et al. 1983).

Some of the recent results from this laboratory on hydrocarbon compositions from hydrothermal
systems are summarized below. Various basic organic geochemical fractionation procedures used
for sample analysis, described in detail elsewhere (e.g., Kawka and Simoneit 1987; Simoneit et al.
1979, 1981, 1984, 1987; Simoneit, Kawka, and Brault 1988; van de Meent et al. 1980; Simoneit
1981, 1982a,b) have been applied here with only minor modifications prior to instrumental analyses
to aid the intercomparability of the resultant data.

Seafloor Spreading Centers and HydrothermalPetroleum Generation

Although this overview discusses the occurrence of petroleum products in eight geographic
locations of hydrothermal activity (figure 16.1), it is my opinion that petroleum generation and
migration is a ubiquitous process associated with hydrothermalism and metallic mineral formation in
global rift systems. All that is really required are organic source material and temperatures in excess
of 50°C in an aqueous medium under a confining pressure. There are at present about 100 known
hydrothermal mineral occurrences at oceanic ridges and rifts (Rona 1988). The quantities of
petroleum associated directly with submarine hydrothermal systems are generally negligible in
terms of economic interest owing to the low sediment overburden, the lack of structural traps, and
often the poor source-rock potential (i.e., low organic carbon content). These systems do, however,
provide a "natural laboratory" for studying the processes of petroleum generation and the behavior
of petroleum in high-temperature fluids. The following data illustrate the diverse suite of
hydrothermal samples from different environments and the wide ranges of organic carbon source
material.

Guaymas Basin, Gulf of California

Guaymas Basin is an actively spreading oceanic basin, which is part of the system of spreading
axes and transform faults that extend from the East Pacific Rise to the San Andreas fault (Curray et
al. 1982; Lonsdale 1985). Ocean-plate accretion occurs by dyke and sill intrusions into the
unconsolidated sediments, leading to high conductive heat flow (Einsele et al. 1980; Curray et al.
1982; Einsele 1985; Lonsdale and Becker 1985). Sediments accumulate rapidly (>2 m/1000 yr) and
have covered the rift floors to a depth of S400 m (Curray et al. 1982). The organic matter of these
recent sediments is derived primarily from diatomaceous and microbial detritus, and the sediments
contain on the average about 2% total organic carbon. Influx of terrigenous organic matter is low
owing to the low continental runoff from the deserts that border the Gulf. Thermal stress causes
rapid maturation with concomitant petroleum generation and expulsion in sediments exposed to
temperatures in excess of about 50°C; the "oil window" seems to migrate upward as the magmatic
heat front rises in the sedimentary column (Simoneit 1984a; Simoneit et al. 1984). Petroleum
products have been characterized in samples obtained by shallow gravity coring in two troughs
(Simoneit et al. 1979), piston coring in the northern trough (Simoneit 1983) and deep coring by Leg
64 of the Deep Sea Drilling Project (DSDP) in the southern trough (Curray et al. 1982). Petroleum-

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (3 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

bearing samples have also been recovered from the seafloor by dredging operations (Simoneit and
Lonsdale 1982) and submersible sampling with the Alvin in 1982, 1985, and 1988 (figure 16.2;
Bazylinski, Farrington, and Jannasch 1988; Simoneit 1984a, b, 1985; Von Damm et al. 1985a;
Kawka and Simoneit 1987; Simoneit and Kawka 1987; Simoneit, Kawka, and Brault 1988).

Leg 64 of the DSDP encountered sills and hydrothermal alteration at depth in all holes drilled in the
basin (figure 16.2). Thermogenic hydrocarbon gas, H2S, and CO2 were identified at all sites. Lipids
(bitumen) were thermally altered close to and below the sills, especially at Site 477 in the southern
trough (Simoneit and Philp 1982; Simoneit et al. 1984). This alteration is indicated by: (1) loss of the
odd carbon number predominance of the n-alkanes (CPI approaches 1.0*); (2) appearance of a
broad hump of UCM (naphthenes); (3) isomerization of various biomarkers; (4) presence of large
amounts of terminal olefins (alk-1-enes), isoprenoid hydrocarbons and elemental sulfur; and (5)
appearance of PAH.

The petroleum products have migrated away from heat sources (e.g., sills or magma at depth) by
advection, diffusion, distillation, and especially by cosolution in hydrothermal fluids (Simoneit and
Philp 1982; Simoneit 1982b; Simoneit et al. 1984). Cosolution in this context refers to the solution in
any proportion of petroleum (bitumen), itself a solution of a large number of compounds, in
hydrothermal water, i.e., the stage before emulsion. The kerogen (i.e., insoluble higher molecular
weight organic matter in sediments) of the shallow DSDP core samples is typical of unaltered
marine organic matter (Simoneit and Philp 1982; Simoneit et al. 1984). In deeper sections (> S180
m depth at Sites 477 and 478), the kerogen reflects essentially complete expulsion of pyrolysate (i.
e., no additional hydrocarbons can be generated by pyrolysis); in these zones of high thermal stress
(entering greenschist facies, 300 50°C, Kastner 1982) the kerogen residues resemble activated
amorphous carbon (Simoneit 1982b), and on the basis of 13C data, this carbon is residual,
thermally matured marine organic carbon (Jenden, Simoneit, and Philp 1982; Simoneit et al. 1984).

Numerous hydrothermal mounds rise to 20-30 m above the southern trough floor (water depth
about 2000 m), and most are actively discharging vent fluids with water temperatures up to 315°C
at S 200 bars, i.e., 0.2 x 108 Pa (Lonsdale 1985; Lonsdale and Becker 1985; Merewether, Olsson,
and Lonsdale 1985). Typical samples from these mounds are stained and cemented with
petroleum; the samples have a strong odor reminiscent of diesel fuel (Simoneit and Lonsdale 1982).
Samples have very diverse petroleum contents and hydrocarbon distributions (figures 16.3, figure
16.4a; Simoneit 1984a, b, 1985; Kawka and Simoneit 1987; Simoneit and Kawka 1987; Simoneit,
Kawka, and Brault 1988) but are similar to those described for bitumens at depth in the DSDP holes
(Simoneit 1983, 1984b). The n-alkanes range from methane to greater than n-C40, with usual
maxima in the mid-C20 region and no carbon number predominance (CPI = 1.0, table 16.1). These
data and the kinetic parameters of the biomarkers indicate that the petroleums were generated by
rapid and intense heating.

An example of a gas chromatographic (GC) trace of an aromatic/naphthenic fraction (F2) of a


sample is shown in figure 16.4b. The major resolved peaks are PAH, a group of compounds
uncommon in petroleums but ubiquitous in higher temperature pyrolysis residues (Geissman, Sim,
and Murdoch 1967; Blumer 1975). The dominant compounds are the pericondensed aromatic
series, for example, phenanthrene, pyrene, benzopyrenes, perylene, benzoperylene, and coronene
(Kawka and Simoneit 1990). A pyrolytic origin is also supported by the presence of five-membered
alicyclic rings (e.g., acenaphthene, methylenephenanthrene, fluorene, fluoranthene, etc.). These
hydrocarbons are found in all pyrolysates from organic matter; once formed, they do not easily

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (4 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

revert to pericondensed aromatic hydrocarbons (Blumer 1975, 1976; Scott 1982). These fractions
also contain significant amounts of toxic PAH, e.g., the benzopyrenes (Kawka and Simoneit 1990).
In addition, perylene is present; it is the predominant PAH of unaltered lipids in sediments deposited
under oxygen-minimum environments in the Gulf (Simoneit 1982a, b; Simoneit and Philp 1982).
Thus, the chemical composition of the aromatic fractions indicates derivation through high-
temperature pyrolysis and rapid quenching, presumably by hydrothermal fluids.

Hydrothermal petroleum migration in Guaymas Basin seems to occur mainly as bulk-phase (pure
petroleum) and to a lesser extent as cosolute fluid and aqueous solution (high-temperature solution
of predominantly n-alkanes only) upward to the seabed, where it condenses (solidifies) and collects
in the conduits and vugs of hydrothermal mineral mounds in response to ambient temperatures.
PAH and sulfur accumulate in the hot vent regions; waxes crystallize in intermediate temperature
areas (20-80°C), and volatile petroleum partly collects in cold areas (3°C) and emanates into the
sea water as plume discharges (Simoneit 1984a,b, 1985; Merewether, Olsson, and Lonsdale 1985;
Simoneit et al. 1990). The activated amorphous carbon residue from the kerogen does not migrate
to the seafloor, although at greater subbottom depths in the DSDP holes this carbon is not evident,
indicating that it may have reacted to oxidative products (e.g., CS2, COS, etc.) or undergone limited
migration (Simoneit 1982b). Both the extensive maturation of organic matter to bitumen in the
DSDP holes and the significant accumulations of petroleum at the rift floor confirm the importance
of hydrothermalism as a feasible mechanism for the formation of petroleum.

Escanaba Trough, Northeastern Pacific

The Escanaba Trough represents the southern extension of the Gorda Ridge, an active oceanic
spreading center about 300 km long, bounded on the north and south by the Blanco and Mendocino
fracture zones, respectively (figure 16.1; McManus et al. 1970). The Trough is filled with up to 500
m of Quaternary turbidite sediments (McManus et al. 1970).

Petroleum cements the sediments and sulfide deposits that blanket the ridge axis and is also
derived from hydrothermal alteration of sedimentary organic matter (Kvenvolden et al. 1986).
Typical GC traces of the saturated and aromatic hydrocarbon fractions are shown in figure .16.5.
The organic source material for these petroleums is mainly terrigenous, on the basis of CPI, carbon-
number range (table 16.1), biomarker composition, and sedimentological considerations
(Kvenvolden and Simoneit 1990).

In the aliphatic hydrocarbon fraction, the n-alkanes range from C14 to C40, with a carbon-number
maximum at n-C27. A predominance of odd carbon numbers > n-C25 (CPI = 1.25, figure 16.5a) is
typical of a contribution from terrestrial, higher plants. Homologs of a marine origin (<n-C25) are
less concentrated. However, these may also have been preferentially biodegraded or expelled by
migration.

The PAH of these samples are mainly the pericondensed nonalkylated series with phenanthrene,
pyrene, and benzopyrenes as the predominant components (figure 16.5b). The overall PAH
compositions are quite variable for these samples, owing probably to the differential solubilities of
the PAH compounds in hot fluids and the temperature ranges of their formation (Simoneit 1984a).
This petroleum was probably generated by intense heating of short duration, as indicated by kinetic
parameters of the biomarkers and by the high concentrations of unsubstituted PAH (figure 16.5b,
Kvenvolden and Simoneit 1990). For example, the aromatic hydrocarbons were interpreted to have

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (5 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

formed at temperatures >400°C when compared with data from laboratory simulations (Kvenvolden
and Simoneit 1990).

Bransfield Strait, Antarctica

Bransfield Strait is a marginal basin located in a heavily sedimented back-arc ridge system (figure
16.1), which is both tectonically and hydrothermally active (Suess et al. 1990). Gravity cores
recovered on the R/V Polarstern cruise ANT IV/2 (1985) were sampled in order to evaluate the
effects of thermal stress on the sedimentary organic matter and to further substantiate the
preliminary data described from a prior cruise (Whiticar, Suess, and Wehner 1985). The general
core lithologies consist of olive-gray clay, which represents thermally unaltered sediment deposited
under generally oxic conditions and black clay/silt sections that are enriched in sulfide, owing
probably to fluid migration. The gray clays near the core bottoms had a slight petroliferous odor, but
migration of heavier hydrocarbons into the sections was not evident.

The major compositional differences of the lipids between the upper and lower sections (subbottom
depths of 5-7 m) of the Bransfield Strait sediments appear to result from an increased rate of
diagenesis in the lower sections (figure 16.6 a,b; Brault and Simoneit 1988). There the temperature
seems to have been higher owing to mild hydrothermal stress, probably from low-temperature fluid
transgression into the sedimentary section. The n-alkane distributions are similar, but the total
concentrations increase slightly with depth. The generation of phytane and other isoprenoids from
phytol via phytenes is evident. Archaebacterial isoprenoids do not appear to be an important input.
Biomarker signature changes in the lower sections include the disappearance of diploptene and the
appearance of hop-17(21)-ene, suggesting diagenetic migration of the double bond.

High abundances of steroidal components in the lipids reflect high productivity and good
preservation (Brault and Simoneit 1988). The 3,5-steradiene (C28) is the dehydration product of
the 5-sterol counterpart. The 2-sterene series is present and seems to have isomerized in the
lower sections to the 4- and 5-sterenes, which are more stable. Diasterenes with the 20R
configuration formed by rearrangement from the sterenes.

Thus, most of the lipid compounds identified at depth in the cores are derived from accelerated
diagenesis of primary biogenic components. The dehydrations of phytol to phytenes and sterols to
sterenes and steradienes are the major indicators of lipid diagenesis with depth, enhanced by mild
hydrothermal stress (Comet, this volume). The minor complex components (e.g., figure 16.6b)
probably represent water-soluble and/or volatile compounds, including gasoline range and aromatic
hydrocarbons (Whiticar, Suess, and Wehner 1985; Whiticar, private communication), brought into
the deeper sections by warm, but not hot, fluid transgression.

East Pacific Rise, 13°N and 21°N

Hydrothermal activity and associated massive sulfide deposits are found on the unsedimented axis
of the East Pacific Rise (EPR) in the region of 13°N (figure 16.1). Abundant faunal communities are
also associated with this activity (Hékinian et al. 1983). Aliphatic hydrocarbons have been analyzed
in hydrothermal plumes and in metalliferous sediments near the active vents and at the base of an
inactive chimney (Brault et al. 1985, 1988). Hydrocarbons from metalliferous sediments have
characteristics of immature organic matter, which was recently biosynthesized and microbiogically

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (6 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

degraded, as indicated by the abundance of low-molecular-weight (C25) n-alkanes and phytane; a


contribution of continental higher plant material is shown by the presence of high-molecular-weight
n-alkanes with an odd carbon number predominance (figure 16.6c, table 16.1). The immature
character of the organic matter is also indicated by the presence of biomarker hydrocarbons derived
from steroids and triterpenoids, which are the result of low-temperature alteration, as might be
expected in the talus of an extinct vent system. The contents of a sediment trap deployed in the
area within S20 m of the vents are characterized by biologically derived material and also by
biomarker compounds affected by thermal stress (Brault et al. 1985), which indicates that higher
temperature degradation of entrained organic detritus is an important process near hydrothermal
discharge sites. Thermally matured compounds (figure 16.6d) are also present at trace levels in
waters collected within S1 km above the hydrothermal vents. The hydrocarbon pattern of these
waters is indicative in many cases of pyrolysis of bacterial matter present in the ocean water that is
entrained during the turbulent cooling of the discharging vent fluids (Brault et al. 1988).

Extensive hydrothermal activity and associated sulfide deposits have also been described at 21°N
on the EPR (figure 16.1) where the crust is unsedimented (Spiess et al. 1980; Ballard et al. 1981;
Von Damm et al. 1985b). The hydrocarbon contents of massive sulfides from vent chimneys are
extremely low (table 16.1) but definitely thermogenic. A GC trace for total hydrocarbons is shown in
figure 16.6e (Brault, Simoneit, and Saliot 1989). The n-alkanes in massive sulfides range from C14
to greater than C40, with no carbon number predominance. The n-alkanes in sulfide with a pyritized
tube worm from a chimney have a slight odd carbon number predominance. All samples contain
PAH, providing evidence for hydrothermal activity. These data, coupled with the carbon-number
maxima at n-C27 or higher, indicate that the hydrocarbons were entrapped/condensed in a high-
temperature regime such as an active chimney. These hydrocarbons were, however, subjected to
these high temperatures for only a short time. The sample with pyritized tube worm residues (figure
16.6e) also contains hydrothermally altered derivatives of biomarkers (e.g., cholestenes, hopenes)
from the vent biota, i.e., mainly tube worms and bacteria. These biomarkers are not present in the
vent biota (Brault and Simoneit, unpublished data).

Atlantis II Deep, Red Sea

The Atlantis II Deep (figure 16.1) contains stratified brine layers, the deepest of which is at a
temperature of 62°C (Hartmann 1980, 1985). Bulk organic matter and hydrocarbons have been
analyzed in two sediment cores (No. 84 and 126, CHAIN 61 cruise) from the Atlantis II Deep
(Simoneit et al. 1987). Although the brine overlying the coring areas is reported to be sterile,
sedimentary organic material derived from autochthonous marine planktonic and microbial inputs
and minor terrestrial sources is present. The organic input derived from the water column above the
brine is further metabolized by microorganisms, and the reworked compounds with organic detritus
are apparently then incorporated into the sediments under the brine by sinking adsorbed or bound
together with particles of metallic oxide precipitates (Simoneit et al. 1987).

Low-temperature maturation in the sediments results in petroleum generation, even from low
amounts of organic matter (average 0.14 wt. percent total organic carbon [TOC]). Both steroid and
triterpenoid hydrocarbons (biomarkers, especially neohop-13(18)-enes) show that extensive acid-
catalyzed reactions are occurring in the sediments (Simoneit et al. 1987). In comparison with other
hydrothermal systems (e.g., Guaymas Basin; Cape Verde Rise, Simoneit et al. 1981), sediments in
the Atlantis II Deep exhibit a lower degree of thermal maturation, as is clearly shown by the
elemental composition (H/C and N/C) of the kerogens and the absence of pyrolytic PAH in the

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (7 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

bitumen.

The lack of carbon-number preference among the n-alkanes (CPI = 1.0) suggests, especially in the
case of the long-chain homologs (e.g., figure 16.7a, table 16.1), that the organic matter has been
affected by catagenesis or that it never had an odd carbon number predominance. However, the
yields of hydrocarbons with respect to sediment weight are much lower than those observed in
other hydrothermal areas. The low temperature and low organic carbon content of the source
sediments in the Atlantis II Deep seem to be responsible for this difference.

Mid-Atlantic Ridge, TAG Area 26°N

The Trans-Atlantic Geotraverse (TAG) hydrothermal field on the Mid-Atlantic Ridge crest at 26°N is
one of two active vent systems known for slow-spreading oceanic ridges (Rona et al. 1984;
Thompson et al. 1988). Hydrothermal deposits lying directly on oceanic crust have been dredged
from the area (TAG 1985-1). Three sulfide-rich samples, consisting mainly of anhydrite, sphalerite,
and chalcopyrite, respectively contained minor amounts of the more volatile (C10-C22)
hydrothermal petroleums (Brault and Simoneit 1989). The saturated and aromatic hydrocarbon
fractions separated from the extract of the sphalerite are shown in figure 16.7 b,c. The n-alkanes
range from C11 to C22 with a CPI = 1.0; pristane and phytane are present, and the UCM maximizes
at the GC retention time for n-C17. This alkane pattern is analogous to that observed for samples
from the EPR at 13°N and is derived from autochthonous marine organic matter. The aromatic
fraction, which contains naphthalene, phenanthrene, their alkyl homologs, and sulfur aromatic
compounds, supports a hydrothermal origin from marine organic matter.

Hydrous Pyrolysis Simulation

Because hydrothermal petroleums are admixtures of products generated over a wide temperature
window, it is important to try to understand the processes through laboratory simulations. A first
attempt using thermally unaltered mud from the seabed of Guaymas Basin (sample 1176-PC2, 10-
20 cm) is described here. The hydrous pyrolysis was carried out in a chromium-lined, stainless steel
vessel with 100 g mud (with 34% pore water) plus 500 g simulated seawater. The temperature
program was 20-330°C over 5 hours, at 330°C for 1 hour, and cooling to 30°C over 2 hours. The
pyrolysis products were extracted from the aqueous and solid phases, separated by liquid
chromatography and analyzed by GC and GC-MS with the conventional procedures.

Homologous Compounds

The n-alkanes generated by the simulation are shown in figure .16.8c and range from C14-C36 with
a minor odd carbon number predominance (CPI = 1.09) and Cmax at C21. This distribution pattern
is completely unlike that of the unaltered sediment (e.g., sample 30G-I, 102-105 cm, figure 16.8a)
but does compare well with that of a hydrothermal petroleum (sample 1172-4, figure 16.8b).
Pristane and phytane are generated with a Pr/Ph of 0.56, which is lower than that for many of the
hydrothermal petroleums, where this ratio is about 1.0. However, the ratios of Pr/n-C17 and Ph/n-
C18 are 0.59 and 0.93, respectively, for the simulation, which more closely approach the same
ratios for the hydrothermal petroleums (e.g., 1.09 and 0.95, respectively, for sample 1172-4). The
naphthenic hump or UCM of the hydrothermal petroleums is generally broad, extending over the full
GC retention range, whereas the lipids of the unaltered sediments exhibit a narrow UCM centered

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (8 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

at about C19 elution time (figure 16.8a, b). The UCM of the simulation bitumen is similar to that
observed for the hydrothermal petroleums.

Most of the hydrothermal petroleums contain a full suite of alkylmonoaromatic hydrocarbons


composed mainly of n-alkylbenzenes, n-alkyltoluenes, and n-alkylxylenes, ranging from C16 to C31
with a Cmax at C23 and no carbon number predominance (figure 16.9a-c). These compounds are
not detectable in the lipids of the unaltered sediment. The simulation bitumen shows n-
alkylbenzenes and n-alkyltoluenes, ranging from C13 to C30 with a lower Cmax at C16 and a slight
odd carbon number predominance (figure 16.9d, e).

Biomarkers

The triterpenoid and steroid hydrocarbon biomarkers are summarized in figure 16.10. The unaltered
sediment contains primarily triterpenes, the 17 (H), 21 (H)-hopane (figure 16.10a) and sterenes ( 2
and 4). A minor amount of the 17 (H), 21 (H)-hopane ( ) series is also present in this repeat
analysis carried out in 1988. The previous GC-MS analysis of 1978 did not exhibit the series
Simoneit et al. 1979; (Kawka and Simoneit 1987). Over the ten years of storage the extract has
changed, losing a major amount of the triterpenes and generating the -hopanes. However, this
does not invalidate the simulation data, because new sample material was used for the
experiments. The mature hydrothermal petroleums contain primarily the -hopane series, ranging
from C27 to C34 (no C28 and traces of C35), with essentially complete maturity for the extended
homologs (C31 22S/R S 1.3, figure 16.10b) (Kawka and Simoneit 1987).

The triterpanes of the simulation experiment are of an intermediate maturity (figure 16.10c). The -
hopanes are predominant (thermally most stable configuration), but a significant amount of the
series remains and a major amount of the intermediate 17 (H),21 (H)-hopane series ( ,
moretanes) is also present. Note that -norhopane is the dominant triterpane in the simulation
sample (C29/C30 = 1.7) as compared with -hopane in the hydrothermal petroleums (C29/C30 =
0.45), and the C29 and C30 moretanes are also abundant in the simulation sample. This
observation appears to indicate a mixture of triterpane biomarkers partially converted to full
maturity. The extended -hopane isomer ratios for C31 and C32 22S/R are 0.3 and 0.5,
respectively, also indicating the immature nature of the triterpanes from the simulation (Ensminger
et al. 1974).

The steranes of the hydrothermal petroleums consist of the C27-C29 homologs ranging in maturity
from dominantly the 5 (H), 14 (H), 17 (H)-20R and lesser amounts of the 5 (H) configuration to
the more rearranged assemblage (e.g., figure .16.10d; Kawka and Simoneit 1987). Diasteranes are
also prominent in the more mature samples (e.g., figure 18.1d). The steranes of the 330°C hydrous
pyrolysis consist mainly of the relatively immature assemblage of the 5 (H), 14 (H), 17 (H)-20R
and the lesser 5 (H) series (figure .16.10e). Diasteranes are not detectable, which may indicate that
clay catalysis by montmorillonite as simulated for Jouy shale (Rubinstein, Sieskind, and Albrecht
1975) is not active in this sediment sample.

Polynuclear Aromatic Hydrocarbons

The hydrothermal petroleums of Guaymas Basin contain PAH consisting primarily of the

file:///F|/Usuarios/Juan%20carlos/articulos/articulo...organic%20matter/Organic%20Matter%20Chapter%2016.htm (9 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

unsubstituted parent compounds (Simoneit 1984a; Kawka and Simoneit 1990), whereas the
unaltered sediments have only perylene from diagenetic sources (Simoneit et al. 1979). The PAH of
the simulation experiment are compared with the hydrothermal oils in figures 16.11 and 16.12.
Phenanthrene and anthracene as well as the alkylphenanthrenes and methylenephenanthrene are
generated in the pyrolysis (figure 16.11d-f) with relative concentrations that are similar but not
identical to typical Guaymas Basin hydrothermal petroleum (figure 16.11a-c). Hydrothermal
petroleums from Escanaba Trough also contain PAH (Kvenvolden and Simoneit 1990), and the
signature of these tricyclic PAH is different (figure 16.11g-i). Anthracene is not detectable and
methylenephenanthrene is more prominent, indicating a higher formation temperature (Kawka and
Simoneit 1990). Diels' hydrocarbon [1,2-(3'-methylcyclopenteno)phenanthrene] was generated in
the simulation in amounts comparable to that found in the Guaymas petroleums. This result
indicates that aromatization of biomarker precursors (e.g., steroids) proceeds rapidly (Simoneit et
al. 1990).

The higher molecular weight PAH are present in the hydrothermal petroleums from Guaymas Basin
and Escanaba Trough (Simoneit 1984a; Kawka and Simoneit 1990; Kvenvolden and Simoneit
1990). For example, the m/z 252 group of benzofluoranthene, benz(e)pyrene, benz(a)pyrene, and
perylene is a significant component in the oils (figure 16.12a-c). Only perylene is a major member of
the m/z 252 series in the oil from the simulation (figure 16.12d), which indicates that higher
temperatures and/or longer heating times are necessary to generate such heavy PAH and that the
generation of perylene from its unknown precursors was enhanced by the conditions of this
simulation.

Discussion

Organic Matter Type

The constitution of the initial organic matter determines the types of petroleum products that form in
basins with a normal geothermal gradient and those with hydrothermal systems. The major source
of petroleum compounds is kerogen, the sedimentary macromolecular organic detritus, which
generally constitutes the bulk of the total organic carbon content (Hunt 1979; Tissot and Welte
1984). In general, terrestrial detritus from mainly vascular plants yields an aromatic kerogen (e.g.,
coal) that has a natural gas potential, whereas marine/lacustrine organic matter from primarily
microbial and planktonic residues yields an aliphatic kerogen (e.g., sapropel) that has a paraffinic
petroleum potential (Hunt 1979; Tissot and Welte 1984). Kerogens in sedimentary basins are
generally mixtures of these inferred endmembers.

Syngenetic sedimentary lipid matter undergoes alteration during diagenetic and early catagenetic
processes that changes the hydrocarbon signature. During oil generation, however, large amounts
of additional hydrocarbons are superimposed on these syngenetic lipids, thus obscuring the earlier
signature (e.g., loss or reduction of the odd carbon number predominance of n-alkanes > C25, cf.
figure 16.8a vs. b). Syngenetic lipids of sediments can usually be utilized to elucidate the various
sources of the total organic matter (e.g., Simoneit 1978, 1981, 1982a). For example, lipid
hydrocarbons of normal sediment in Guaymas Basin (figure 16.8a) contain n-alkanes, mostly <
C21, and a minor series of homologs > C23 with an odd carbon number predominance. The lower
carbon number composition is typical of a predominantly marine planktonic and microbial origin with
a minor influx of terrigenous vascular plant wax (> C25). By contrast, lipid hydrocarbons of sediment

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (10 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

in the Escanaba Trough (Kvenvolden et al. 1986; Simoneit 1988) contain n-alkanes, mostly > C23
with an odd carbon number predominance. This signature is probably derived mainly from
terrigenous vascular plant wax, with a minor microbial component (Kvenvolden et al. 1986). In both
the Guaymas and Escanaba sediment samples, the compositions of the kerogens support these
interpretations based on the nature of the lipid hydrocarbons (Simoneit 1978; Simoneit et al. 1979).

Petroleum Generation

The principal zone of petroleum formation in sedimentary sequences under normal geothermal
gradients extends from about 1 to 3 km depth (e.g., Hunt 1979; Tissot and Welte 1984). This depth
corresponds to a temperature range of 50°-120°C. The effect of pressure on this process has not
been quantified (Tissot and Welte 1984). The further cracking of organic matter to natural gas is
thought to take place at higher temperatures of 150°-250°C, although gas is also generated within
the "oil window" (e.g., Kartsev et al. 1971; Vassoevich, Akramkhodzhaev, and Geodekyan 1974;
Hunt 1979). These proposed temperature regimes for the oil and gas "windows" need some upward
adjustment to accommodate the recent data on hydrocarbon presence at higher temperatures in
hydrothermal systems and ultradeep wells.

The "instantaneous" petroleum generation in hydrothermal systems is a natural and efficient


process, which occurs at temperatures approaching a maximum of 400°C. At such high
temperatures, organic matter is only partly destroyed, probably because the thermogenic products
are rapidly removed from the hot zone. Formation of hydrothermal petroleum seems to commence
in low-temperature areas, generating the more aliphatic products (Simoneit, Kawka, and Brault
1988), and as the temperature regime rises, products are derived from more refractory organic
matter and may even be "resynthesized" (e.g., PAH compounds).

For example, the phenanthrene-to-methylphenanthrenes (P/MP) ratio can be used to estimate


formation temperatures for these compounds by comparison with laboratory data (figure 16.11).
The P/MP for the Guaymas Basin oils range from 0.12 to 0.95, with an average of 0.58 (excluding a
low and high outlier), and for the Escanaba Trough oils, from 2.7 to 3.3 (Kawka and Simoneit 1990;
Kvenvolden and Simoneit 1990). These ratios can be compared with data from laboratory
maturation studies of immature Tanner Basin (California Bight) kerogen (Ishiwatari and Fukushima
1979), a process analogous to hydrothermal organic matter alteration. The phenanthrene-to-
methylphenanthrene ratios ranged from 0.59-0.83 at 310°C for 18 to 100 hours, and 0.9-5.0 at 410°
C for 5 to 32 hours. Based on this ratio, the GB oils fit with a temperature window of around 310°C
and the ET oils with temperatures approaching 410°C.

Organic matter associated with deeper hydrothermal systems (e.g., epithermal ores in volcanic
terranes) is typically more asphaltic and has a high PAH content. Such organic matter is widely
distributed (e.g., California mercury deposits: Geissman, Sim, and Murdoch 1967; Blumer 1975;
other hydrothermal sulfides: Germanov and Bannikova 1972). Deep well drilling (>7000 m) has
intersected Cretaceous shales that were at in situ temperatures of about 260°-300°C (e.g., Price,
Clayton, and Rumen 1981; Price 1982). These samples were rich in bitumen components and their
kerogens still had significant hydrocarbon generation potential. These observations indicate that in
situ petroleum is stable at much higher temperatures and pressures than those discussed above
and probably also over longer geologic time periods.

The major similarities and differences between hydrothermal petroleums and reservoir petroleums

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (11 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

have been summarized (Simoneit 1988) and are briefly recapped here. table 16.2 presents this
summary in terms of compound classes and bulk parameters. Most hydrocarbon products occur in
both types of petroleum; a major difference is the enhanced content of PAH and sulfur-PAH in the
hydrothermal products.

Migration Processes

Petroleum migration in conventional sedimentary basins is proposed to occur mainly by oil phase
flow, and migration in water solution and/or by molecular diffusion is of minor importance (Durand
1988). However, in hydrothermal systems the migration of petroleum occurs by a different
combination of the processes mentioned above, and the following laboratory measurements on
petroleum solubility in water (Price 1976; Price et al. 1983) may be applicable. The aqueous
solubility of petroleums and various hydrocarbon fractions increases exponentially from 100°C to
180°C, and solubilities are high enough to possibly account for the formation of petroleum
reservoirs by migration of molecular or cosolutions (Price 1976). Increased salinities of 150 NaCl
cause dramatic exsolution of the petroleum and at 350 almost total "salt-out" occurs (Price 1976).
This finding is consistent with the requirement for the separation of petroleum from migrating
solutions in the salty waters of reservoir sands. It has been demonstrated that methane in the
presence of water is an even better carrier for petroleum than water or methane alone (Price et al.
1983). Both increases in pressure (to about 1800 bar, i.e., 1.8 x 108 Pa) and in temperature (to 250°
C) raised the solubility of petroleum. Cosolubility was found at rather moderate conditions (e.g., 100°
C at 108 Pa, 200°C at 0.5 x 108 Pa). The addition of other gases such as CO2 and ethane to this
mixture also increases the solubility of petroleum. Under these experimental conditions, which are
similar to those in hydrothermal systems, CH4, C2H6, and CO2 are all supercritical, and H2O
approaches its near-critical state; thus, the gases and H2O are all mutually soluble by the reduced
hydrogen bonding in water and are a good cosolvent for petroleum. Therefore, in hydrothermal
systems petroleum expulsion and primary migration appear to proceed as gas/fluid and aqueous
solution phases and secondary migration as bulk phase and to a lesser extent in cosolution,
emulsion, and by molecular diffusion (Simoneit 1985, 1988; Kawka and Simoneit 1987; Simoneit,
Kawka and Brault 1988; Didyk and Simoneit 1989; Simoneit et al. 1990).

Simulation

The results of the laboratory hydrothermal alteration of Guaymas Basin sediment are summarized
in table 16.3. Generally, most of the compounds found in the hydrothermal petroleums were
generated by the simulation experiment. This has also been observed for other hydrous pyrolysis
experiments where the match of crude oil to pyrolysis products from the source rock is essentially
identical (e.g., Lewan 1985; Lewan, Bjorøy, and Dolcater 1986). Thus, the bulk of the hydrothermal
petroleum compounds (i.e., n-alkanes, naphthenes, isoprenoids) form readily over a range
commencing at low temperatures to the high temperatures of this simulation. The hopane and
sterane biomarkers were not converted to the fully mature configurations, and sterane maturation
was slower than hopane maturation. This is probably due to the brief heating time and is as
expected from other studies.

The alkylarenes appeared at the onset of generation in the simulation, and the lower molecular
weight PAH, such as for example, the phenanthrene series, are also fully represented. Triaromatic
steroid hydrocarbons, including Diels' hydrocarbon, were generated in the hydrous pyrolysis.
However, the higher molecular weight PAH, excluding perylene, were not generated in the

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (12 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

simulation. This indicates that aromatization of steroids and other tricyclic precursors proceeds
readily, but higher temperatures, longer contact times, and/or unknown reactions are required to
form the heavy PAH.

Recent immature sediments receive biogenic detritus, which upon deposition, undergoes diagenetic
and additional microbial alteration. Increasing burial in sedimentary basins results in the onset of
organic matter maturation, which generates some volatile products from the kerogen (easily
cracked moieties) that become added to the endogenous lipid residues. This is the beginning of
petroleum formation. As the depth of burial (i.e., temperature) increases, catagenesis commences
and major petroleum generation takes place. At still greater depths of burial the metagenetic stage
is reached, where extensive cracking, disproportionation, and reforming of the organic matter, both
petroleum and kerogen residues, occur to yield primarily gases and residual carbon.

In the case of hydrothermal systems, these processes are compressed into an "instantaneous"
geological time frame. At seafloor spreading axes, hydrothermal systems operating below a
sediment blanket (e.g., Guaymas Basin and Escanaba Trough) generate petroleum from generally
immature organic matter in the sediments. This petroleum then migrates upward, leaving behind a
spent carbonaceous residue (amorphous "activated" carbon). The Guaymas petroleums are
composed of: (1) gasoline-range hydrocarbons (C1-C12); (2) a broad distribution of n-alkanes (C12-
C40+) with no carbon number predominance; (3) a naphthenic hump, UCM; (4) pristane and
phytane at significant concentrations; (5) mature biomarkers (e.g., -hopanes); and (6) significant
concentrations of PAH and sulfur. Exposed petroleum and petroleums in unconsolidated surface
sediments are microbially degraded and leached, whereas interior samples are essentially
unaltered.

Hydrothermal systems operating in unsedimented rift areas (e.g., East Pacific Rise at 13°N and 21°
N, Mid-Atlantic Ridge at 26°N) generate trace amounts of petroleum. Low amounts of petroleum are
generated by pyrolysis of suspended and dissolved biogenic organic detritus (including bacteria and
algae) entrained during the turbulent cooling of the vents and in the talus for both sedimented and
bare rock systems. However, this type of petroleum is overwhelmed in sedimented systems by the
large quantity of petroleum generated from the sedimentary organic matter. In addition, low-level
maturation is observed in the surrounding area at vent sites, owing probably to warming of ambient
detritus.

The preliminary laboratory simulation (330°C hydrous pyrolysis) of sediment alteration indicates that
the products are intermediate in terms of maturity between source sediment and hydrothermal
petroleum. The duration of heating was too brief to convert all precursors to their end products. The
major simulation products are the same as in the hydrothermal oils. However, the biomarkers are
not as mature and heavy PAH have not yet formed. The unaltered sediment contains immature
lipids and no PAH, except for some perylene. Thus, hydrothermal petroleums at the seabed in
active rift zones under a sedimentary cover are accumulated mixtures of low- to high-temperature
hydrous pyrolysis products.

Acknowledgments

I thank the Deep Sea Drilling Project for access to DSDP samples, and the crews and pilots of the
D.S.V. Alvin, R. V. Lulu, and R. V. Atlantis II for their skillful recoveries of hydrothermal samples.
Samples, data, and assistance were provided by C. A. Allen, J. Baross, M. Brault, A. Grey, P. D.

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (13 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

Jenden, O. E. Kawka, K. A. Kvenvolden, R. N. Leif, P. F. Lonsdale, A. Lorre, M. A. Mazurek, R. P.


Philp, P. A. Rona, E. Ruth, G. Stormberg, E. Suess, and G. M. Wang. The hydrous pyrolysis
experimentation was carried out in collaboration with the Idaho National Engineering Laboratory,
EG&G Idaho, Inc., Idaho Falls, Idaho. Funding from the Division of Ocean Sciences, National
Science Foundation (Grants OCE81-18897, OCE-8312036, OCE-8512832, and OCE-8601316) is
gratefully acknowledged.

References

Ballard, R. D., J. Francheteau, T. Juteau, C. Rangan, and W. Normark. 1981. East Pacific
Rise at 21°N: The volcanic, tectonic and hydrothermal processes of the central axis. Earth
Planet. Sci. Lett. 55:1-10.

Bazylinski, D. A., J. W. Farrington, and H. W. Jannasch. 1988. Hydrocarbons in surface


sediments from a Guaymas Basin hydrothermal vent site. Org. Geochem. 12:547-558.

Blumer, M. 1975. Curtisite, idrialite and pendletonite, polycyclic aromatic hydrocarbon


minerals: their composition and origin. Chem. Geol. 16:245-256.

Blumer, M. 1976. Polycyclic aromatic compounds in nature. Scientific American 234(3):34-


45.

Brault, M. and B. R. T. Simoneit. 1988. Steroid and terpenoid distributions in Bransfield Strait
sediments: hydrothermally enhanced diagenetic transformations. In Advances in Organic
Geochemistry 1987. Org. Geochem. 13:697-705.

Brault, M. and B. R. T. Simoneit. 1989. Trace petroliferous organic matter associated with
hydrothermal minerals from the Mid-Atlantic Ridge at the Trans-Atlantic Geotraverse 26°N
Site. J. Geophys. Res. 94:9791-9798.

Brault, M., B. R. T. Simoneit, and A. Saliot. 1989. Trace petroliferous organic matter
associated with massive hydrothermal sulfides from the East Pacific Rise at 13°N and 21°N.
Oceanologica Acta 12:405-415.

Brault, M., B. R. T. Simoneit, J. C. Marty, and A. Saliot. 1985. Les hydrocarbures dans le
systeme hydrothermal de la ride Est-Pacifique, a 13°N. C.R. Acad. Sci. Paris 301, (II):807-
812.

Brault, M., B. R. T. Simoneit, J. C. Marty, and A. Saliot. 1988. Hydrocarbons in waters and
particulate material from hydrothermal environments at the East Pacific Rise, 13°N. Org.
Geochem. 12:209-219.

Cooper, J. E. and E. E. Bray. 1963. A postulated role of fatty acids in petroleum formation.
Geochim. Cosmochim. Acta 29:1113-1127.

Curray, J. R., D. G. Moore, et al. 1982. Initial Reports. DSDP, 64, parts 1 and 2. Washington,
D.C.: U.S. Govt. Printing Office.

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (14 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

Demaison, G. J. and G. T. Moore. 1980. Anoxic environments and oil source bed genesis.
Org. Geochem. 2:9-31.

Didyk, B. M. and B. R. T. Simoneit. 1989. Hydrothermal oil of Guaymas Basin and


implications for petroleum formation mechanisms. Nature 342:65-69.

Didyk, B. M., B. R. T. Simoneit, S. C. Brassell, and G. Eglinton. 1978. Organic geochemical


indicators of paleoenvironmental conditions of sedimentation. Nature 272:216-222.

Durand, B. 1988. Understanding of HC migration in sedimentary basins (present state of


knowledge). In L. Mattavelli and L. Novelli, eds., Advances in Organic Geochemistry 1987.
Org. Geochem. 13:445-459.

Einsele, G. 1985. Basaltic sill-sediment complexes in young spreading centers: genesis and
significance. Geology 13:249-252.

Einsele, G., J. Gieskes, J. Curray, D. Moore, E. Aguayo, M. P. Aubry, D. J. Fornari, J. C.


Guerrero, M. Kastner, K. Kelts, M. Lyle, Y. Matoba, A. Molina-Cruz, J. Niemitz, J. Rueda, A.
Saunders, H. Schrader, B. R. T. Simoneit, and V. Vacquier. 1980. Intrusion of basaltic sills
into highly porous sediments and resulting hydrothermal activity. Nature 283:441-445.

Ensminger, A., A. van Dorsselaer, C. Spyckerelle, P. Albrecht, and G. Ourisson. 1974.


Pentacyclic triterpanes of the hopane type as ubiquitous geochemical markers: Origin and
significance. In B. Tissot and F. Bienner, eds., Advances in Organic Geochemistry 1973, pp.
245-260, Paris: Technip.

Geissman, T. A., K. Y. Sim, and J. Murdoch. 1967. Organic minerals. Picene and chrysene
as constituents of the mineral curtisite (idrialite). Experientia 23:793-794.

Germanov, A. I. and L. A. Bannikova. 1972. Alteration of organic matter of sedimentary rocks


during hydrothermal sulfide concentration. Dokl. Akad. Nauk SSSR 203:1180-1182 (in
Russian).

Hartmann, M. 1980. Atlantis II Deep geothermal brine system. Hydrographic situation in


1977 and changes since 1965. Deep Sea Res. 27:161-171.

Hartman, M. 1985. Atlantis II Deep geothermal brine system. Chemical processes between
hydrothermal brines and Red Sea deep water. Mar. Geol. 64:157-177.

Hékinian, R., M. Fevrier, F. Avedik, P. Cambon, J. L. Charlou, H. D. Needham, J. Raillard, J.


Boulegue, L. Merlivat, A. Moinet, S. Manganini, and J. Lange. 1983. East Pacific Rise near
13°N: Geology of new hydrothermal fields. Science 219:1321-1324.

Hunt, J. M. 1979. Petroleum Geochemistry and Geology. San Francisco: W. H. Freeman.

Ishiwatari, R. and K. Fukushima. 1979. Generation of unsaturated and aromatic

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (15 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

hydrocarbons by thermal alteration of young kerogen. Geochim. Cosmochim. Acta 43:1343-


1349.

Jenden, P. D., B. R. T. Simoneit, and R. P. Philp. 1982. Hydrothermal effects on


protokerogen of unconsolidated sediments from Guaymas Basin, Gulf of California:
Elemental compositions, stable carbon isotope ratios and electronspin resonance spectra. In
J. R. Curray, D. G. Moore, et al., eds., Initial Reports. DSDP 64:905-912. Washington, D.C.:
U.S. Govt. Printing Office.

Johns, R. B., ed. 1986. Biological Markers in the Sedimentary Record, Methods in
Geochemistry and Geophysics 24. Amsterdam: Elsevier.

Kartsev, A. A., N. B. Vassoevich, A. A. Geodekian, S. G. Neruchev and V. A. Sokolov. 1971.


The principal stage in formation of petroleum. Proc. 8th World Petrol. Congr. 2:3-11.

Kastner, M. 1982. Evidence for two distinct hydrothermal systems in the Guaymas Basin. In
J. R. Curray, D. G. Moore, et al., eds., Initial Reports. DSDP 64:1143-1157. Washington, D.
C.: U.S. Govt. Printing Office.

Kawka, O. E. and B. R. T. Simoneit. 1987. Survey of hydrothermally generated petroleums


from the Guaymas Basin spreading center. Org. Geochem. 11:311-328.

Kawka, O. E. and B. R. T. Simoneit. 1990. Polycyclic aromatic hydrocarbons in hydrothermal


petroleums from the Guaymas Basin spreading center. In B. R. T. Simoneit, ed., Organic
Matter Alteration in Hydrothermal Systems-Petroleum Generation, Migration and
Biogeochemistry. Appl. Geochem. 5:17-27.

Kvenvolden, K. A. and B. R. T. Simoneit. 1990. Hydrothermally derived petroleum: examples


from Guaymas Basin, Gulf of California and Escanaba Trough, Northeast Pacific Ocean.
AAPG Bull. 74:223-237.

Kvenvolden, K. A., J. B. Rapp, F. D. Hostettler, J. L. Morton, J. D. King, and G. E. Claypool.


1986. Petroleum associated with polymetallic sulfide in sediment from Gorda Ridge. Science
234:1231-1234.

Lewan, M. D. 1985. Evaluation of petroleum generation by hydrous pyrolysis


experimentation. Phil. Trans. Roy. Soc. Lond. A315:123-134.

Lewan, M. D., M. Bjorøy, and D. L. Dolcater. 1986. Effects of thermal maturation on steroid
hydrocarbons as determined by hydrous pyrolysis of Phosphoria Retort Shale. Geochim.
Cosmochim. Acta 50:1977-1987.

Lonsdale, P. 1985. A transform continental margin rich in hydrocarbons, Gulf of California.


AAPG Bull. 69:1160-1180.

Lonsdale, P. and K. Becker. 1985. Hydrothermal plumes, hot springs, and conductive heat
flow in the Southern Trough of Guaymas Basin. Earth Planet. Sci. Lett. 73:211-225.

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (16 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

McManus, D. A., R. E. Burns, et al. 1970. Initial Reports of the Deep Sea Drilling Project, 5.
Washington, D.C.: U.S. Govt. Printing Office.

Merewether, R., M. S. Olsson, and P. Lonsdale. 1985. Acoustically detected hydrocarbon


plumes rising from 2-km depths in Guaymas Basin, Gulf of California. J. Geophys. Res.
90:3075-3085.

Price, L. C. 1976. Aqueous solubility of petroleum as applied to its origin and primary
migration. AAPG Bull. 60:213-244.

Price, L. C. 1982. Organic geochemistry of core samples from an ultra-deep hot well (300°C,
7 km). Chem. Geol. 37:215-228.

Price, L. C., J. S. Clayton, and L. L. Rumen. 1981. Organic geochemistry of the 9.6 km
Bertha Rogers No. 1 well, Oklahoma. Org. Geochem. 3:59-77.

Price, L. C., L. M. Wenger, T. Ging, and C. W. Blount. 1983. Solubility of crude oil in
methane as a function of pressure and temperature. Org. Geochem. 4:201-221.

Rona, P. A. 1988. Hydrothermal mineralization at oceanic ridges. In T. J. Barrett and J. L.


Jambor, eds., Seafloor Hydrothermal Mineralization. Min. Assoc. Canada. Can. Min. 26:431-
465.

Rona, P. A., G. Thompson, M. J. Mottl, J. A. Karson, W. J. Jenkins, D. Graham, M. Mallette,


K. von Damm, and J. M. Edmond. 1984. Hydrothermal activity at the Trans-Atlantic
Geotraverse hydrothermal field, Mid-Atlantic Ridge Crest at 26°N. J. Geophys. Res.
89:11365-11377.

Rubinstein, I., O. Sieskind, and P. Albrecht. 1975. Rearranged sterenes in a shale:


occurrence and simulated formation. J. Chem. Soc., Perk. Trans. 1:1833-1876.

Scott, L. T. 1982. Thermal rearrangements of aromatic compounds. Acc. Chem. Res. 15:52-
58.

Simoneit, B. R. T. 1978. The organic chemistry of marine sediments. In J. P. Riley and R.


Chester, eds., Chemical Oceanography 7:233-311. London, Academic.

Simoneit, B. R. T. 1981. Utility of molecular markers and stable isotope compositions in the
evaluation of sources and diagenesis of organic matter in the geosphere. In A. Prashnowsky,
ed., The Impact of the Treibs' Porphyrin Concept on the Modern Organic Geochemistry, pp.
133-158, Würzburg: Bayerische Julius Maximilian Universität.

Simoneit, B. R. T. 1982a. The composition, sources and transport of organic matter to


marine sediments-the organic geochemical approach. In J. A. J. Thompson and W. D.
Jamieson, eds., Proceedings of the Symposium on Marine Chemistry into the Eighties, pp.
82-112, Natl. Res. Council of Canada.

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (17 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

Simoneit, B. R. T. 1982b. Shipboard organic geochemistry and safety monitoring, Leg. 64,
Gulf of California. In J. R. Curray, D. G. Moore, et al., eds., Initial Reports. DSDP 64:723-
728. Washington, D.C.: U.S. Govt. Printing Office.

Simoneit, B. R. T. 1983. Organic matter maturation and petroleum genesis: geothermal


versus hydrothermal. In The Role of Heat in the Development of Energy and Mineral
Resources in the Northern Basin and Range Province, pp. 215-241. Davis, Calif.: Geotherm.
Res. Council, Special Report no. 13.

Simoneit, B. R. T. 1984a. Hydrothermal effects on organic matter-high versus low


temperature components. In Advances in Organic Geochemistry 1983. Org. Geochem.
6:857-864.

Simoneit, B. R. T. 1984b. Effects of hydrothermal activity on sedimentary organic matter:


Guaymas Basin, Gulf of California-petroleum genesis and protokerogen degradation. In P. A.
Rona, K. Boström, L. Laubier, and K. L. Smith, Jr., eds., Hydrothermal Processes at Seafloor
Spreading Centers, pp. 453-474. New York: NATO-ARI Series, Plenum.

Simoneit, B. R. T. 1985. Hydrothermal petroleum: genesis, migration and deposition in


Guaymas Basin, Gulf of California. Can. J. Earth Sci. 22:1919-1929.

Simoneit, B. R. T. 1988. Petroleum generation in submarine hydrothermal systems: An


update, Can. Mineralogist 26:827-840.

Simoneit, B. R. T., ed. 1990. Organic matter alteration in hydrothermal systems-petroleum


generation, migration and biogeochemistry. Appl. Geochem. 5:1-248.

Simoneit, B. R. T. and O. E. Kawka. 1987. Hydrothermal petroleum from diatomites in the


Gulf of California. In J. Brooks and A. J. Fleet, eds., Marine Petroleum Source Rocks, pp.
217-228, Geol. Soc. Lond., spec. publ. no. 26.

Simoneit, B. R. T. and P. F. Lonsdale. 1982. Hydrothermal petroleum in mineralized mounds


at the seabed of Guaymas Basin. Nature 295:198-202.

Simoneit, B. R. T. and R. P. Philp. 1982. Organic geochemistry of lipids and kerogen and the
effects of basalt intrusions on unconsolidated oceanic sediments: Sites 477, 478 and 481,
Guaymas Basin, Gulf of California. In J. R. Curray, D. G. Moore, et al., eds., Initial Reports.
DSDP 64:883-904. Washington, D.C.: U.S. Govt. Printing Office.

Simoneit, B. R. T., O. E. Kawka, and M. Brault. 1988. Origin of gases and condensates in
the Guaymas Basin hydrothermal system, (Gulf of California). In M. Schoell, ed. Proc. Symp.
Origins of Methane in the Earth. Chem. Geol. 71:169-182.

Simoneit, B. R. T., S. Brenner, K. E. Peters, and I. R. Kaplan. 1981. Thermal alteration of


Cretaceous black shale by basaltic intrusions in the eastern Atlantic. II: Effects on bitumen
and kerogen. Geochim. Cosmochim. Acta 45:1581-1602.

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (18 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

Simoneit, B. R. T., J. O. Grimalt, J. M. Hayes, and H. Hartman. 1987. Low temperature


hydrothermal maturation of organic matter in sediments from the Atlantis II Deep, Red Sea.
Geochim. Cosmochim. Acta 51:879-894.

Simoneit, B. R. T., P. F. Lonsdale, J. M. Edmond, and W. C. Shanks, III. 1990. Deep-water


hydrocarbon seeps in Guaymas Basin, Gulf of California. In B. R. T. Simoneit, ed., Organic
Matter Alteration in Hydrothermal Systems-Petroleum Generation, Migration and
Biogeochemistry. Appl. Geochem., 5:41-49.

Simoneit, B. R. T., R. P. Philp, P. D. Jenden, and E. M. Galimov. 1984. Organic


geochemistry of Deep Sea Drilling Project sediments from the Gulf of California
hydrothermal effects on unconsolidated diatom ooze. Org. Geochem. 7:173-205.

Simoneit, B. R. T., M. A. Mazurek, S. Brenner, P. T. Crisp and I. R. Kaplan. 1979. Organic


geochemistry of recent sediments from Guaymas Basin, Gulf of California. Deep Sea Res.
26A:879-891.

Spiess, F. N., K. C. Macdonald, T. Atwater, E. Ballard, A. Carranza, D. Cordoba, C. Cox, V.


M. DiazGarcia, J. Francheteau, J. Guerrero, J. Hawkins, R. Haymon, R. Hessler, T. Juteau,
M. Kastner, R. Larson, B. Luyendyk, J. D. Macdougall, S. Miller, W. Normark, J. Orcutt, and
C. Rangin. 1980. East Pacific Rise; hot springs and geophysical experiments. Science
207:1421-1433.

Suess, E., B. R. T. Simoneit, G. Wefer, M. J. Whiticar, M. Fisk, M. von Breymann, M. W.


Han, R. Wittstock, C. Laban, D. Kadko, and Z. Top. 1990. Hydrothermalism in the Bransfield
Strait, Antarctica. Geologische Rundschau, in preparation.

Thompson, G., S. E. Humphris, B. Schroeder, M. Sulanowska, and P. A. Rona. 1988.


Hydrothermal mineralization on the Mid-Atlantic Ridge. Can. Min. 26:697-711.

Tissot, B. P. and D. H. Welte. 1984. Petroleum Formation and Occurrence: A New Approach
to Oil and Gas Exploration. Berlin: Springer Verlag.

van de Meent, D., S. C. Brown, R. P. Philp, and B. R. T. Simoneit. 1980. Pyrolysis-high


resolution gas chromatography and pyrolysis gas chromatography-mass spectrometry of
kerogens and kerogen precursors. Geochim. Cosmochim. Acta 44:999-1013.

Vassoevich, N. B., A. M. Akramkhodzhaev, and A. A. Geodekyan. 1974. Principal zone of oil


formation. In B. Tissot and F. Bienner, eds., Advances in Organic Geochemistry 1973, pp.
309-314. Paris: Technip.

Von Damm, K. L., J. M. Edmond, C. I. Measures, and B. Grant. 1985a. Chemistry of


submarine hydrothermal solutions at Guaymas Basin, Gulf of California. Geochim.
Cosmochim. Acta 49:2221-2237.

Von Damm, K. L., J. M. Edmond, B. Grant, C. I. Measures, B. Walden, and R. F. Weiss.


1985b. Chemistry of submarine hydrothermal solutions at 21°N, East Pacific Rise. Geochim.

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (19 de 20)17/01/2006 06:48:20 p.m.


Organic Matter: Chapter 16

Cosmochim. Acta 49:2197-2220.

Whiticar, M. J., E. Suess, and H. Wehner. 1985. Thermogenic hydrocarbons in the surface
sediments of the Bransfield Strait, Antarctic Peninsula. Nature 314:87-90.

*CPI=carbon preference index: for hydrocarbons it is expressed as a summation of the odd


carbon number homologs over a defined range, divided by a summation of the even carbon
number homologs over the same range (Cooper and Bray 1963; Simoneit 1978).

Organic Matter: Productivity, Accumulation, and Preservation in Recent and Ancient Sediments

file:///F|/Usuarios/Juan%20carlos/articulos/articul...rganic%20matter/Organic%20Matter%20Chapter%2016.htm (20 de 20)17/01/2006 06:48:20 p.m.

Potrebbero piacerti anche