You are on page 1of 17

Controls on Hydrogen Sulfide

Formation in a Jurassic Carbonate


Play, Turkmenistan
Gary H. Isaksen
ExxonMobil Exploration Company, Houston, Texas, U.S.A.
Mukhammetnur Khalylov
State Concern Turkmengas, Institute of Oil and Gas, Ashgabat, Turkmenistan
ABSTRACT
A
n integrated geological and geochemical evaluation of the Charjou terrace
area of the Amu Darya Basin in Turkmenistan and Uzbekistan shows
that the main controls on hydrogen sulfide (H
2
S) distribution are the
lithofacies of the reservoir and seal rocks and reservoir temperature. Thermo-
chemical sulfate reduction (TSR) has likely occurred where reservoir tempera-
tures are greater than approximately 1008C, and the reservoired gas is trapped in
dolomitic reservoirs in contact with anhydrite. These chemical reactions occur
both in carbonate platform and reefal buildups capped by anhydrite, as well as
in structural traps that have undergone thrust-induced juxtapositioning of an-
hydrite with the dolomite gas reservoir. The reservoirs can have an H
2
S con-
centration as much as 4 vol.%. This, together with temperatures of 1001108C
and sulfate-rich formation waters, suggests that these reactions represent the
earliest stage of TSR. Carbonate buildups, including reefs, capped by marine shales
have very lowto trace levels of H
2
S because the thermochemical sulfate reaction is
not initiated, presumably caused by a lack of sulfate. Furthermore, H
2
S concen-
trations can vary greatly within a single field with stacked pay zones. Reservoir
compartments sealed by marine shales have no H
2
S gas. A covariance between
increased gas wetness and lower initial temperatures for TSR is also suggested.
INTRODUCTION
Turkmenistans oil and gas reserves are substantial,
bothinthe context of the Caspianregionas well as on
a worldwide scale. In a recent review of the countrys
oil and gas resources, Dorian (2002) references U.S.
Energy Information Administration data for February
2002 of a (highside) estimated22.8 trillionm
3
(805 tcf)
of gas and 12 billion tons (88 billion bbl) of oil Belo-
polsky and Talwani (2007) quote proven gas reserves
of 200 tcf for the Amu Darya Basin. In comparison,
Azerbaijans resources are estimated at 46 tcf gas and
Chapter 14
Isaksen, G. H., and M. Khalylov, 2007, Controls on hydrogen sulfide
formation in a Jurassic carbonate play, Turkmenistan, in P. O. Yilmaz
and G. H. Isaksen, editors, Oil and gas of the Greater Caspian area: AAPG
Studies in Geology 55, p. 133149.
133
Copyright n2007 by The American Association of Petroleum Geologists.
DOI:10.1306/1205843St551436
3645 billion bbl of oil, whereas Kazakhstans gas re-
sources are estimatedat 153158tcf, andoil resources
are estimated at 102110 billion bbl. The bulk of Turk-
menistans gas reserves are located in its eastern ba-
sins, which account for 80% of the countrys annual
gas production. The largest among the near 70 gas
and gas-condensate discoveries is at the Douletabad
Donmez and Shatlyk complexes in the south and
ChartakMalai, Semanteppe, Beshgyzyl, and Bota
fields in the northeast. Since the late 1990s, com-
mercial flowrates of oil have been made at the Eloten
and Seyrap fields inthe central Amu Darya Basin. The
gas is produced by State Concern Turkmengas.
Gas quality, most importantlyhydrogensulfide (H
2
S)
content, is of particular concern in this basin. Occur-
rence of H
2
S in natural gas is of great economic impor-
tance because large monetary investments are required
during development and production and also to build
so-called scrubbing facilities for gas processing, instal-
lations that commonly use an alkanolamine or iron-
sponge sweetening process to remove H
2
S from natu-
ral gas. Amine processing has the added advantage of
being able to remove CO
2
. The presence of H
2
S in the
production or transport stream causes corrosion or
sulfide stress cracking of carbon and low-alloy steels
and enhances chloride stress corrosion of high-alloy
(stainless) steels. Naturally, the toxicity of H
2
S is also a
major concern for environment, health, and safety.
A gas is classified as sour when H
2
S concentrations
exceed 4 ppm(0.25 g/100 ft
3
) (Hyne, 1991). H
2
S con-
centrations in natural gases in the Charjou area of
Turkmenistanrange upward to 4%volume. Basins else-
where in the world are known to contain sour gas in
far greater concentrations. The area north of the Wig-
gens archinthe Smackover trendoffshore Mississippi,
U.S.A., contains gas accumulations with 77% H
2
S
(Rhodes, 1994). Similar values were reportedby Wade
et al. (1989) andHeydari (1997) for the Smackover trend
onshore Mississippi, Alabama, andFlorida. Other exam-
ples of sour gas provinces include the Middle and Up-
per Devonian (Leduc, Beaverhill Lake, and Nisku for-
mations) carbonates of south-central Alberta (Canada)
(Krouse et al., 1988; Manzano et al., 1997); the Permian
Khuff Formation of Abu Dhabi (Worden and Smalley,
1996); the Permian KupferschieferZechstein of Ger-
many (Bechtel and Puettmann, 1991); and the Upper
Jurassic and Cretaceous reefs of the Lacq field in Aqui-
taine, France (ConnanandLacrampe-Couloume, 1993).
The occurrence of H
2
S in natural gas is typically
associated with marine, carbonate hydrocarbon sys-
tems. In such environments, H
2
S can be formed by a
variety of processes, including direct generationfrom
thermal maturation of sulfur-rich (type II-S) source
rocks and/or sulfate reductionby bacterial or thermo-
chemical means.
This study is concerned with the hydrocarbon
systemin the northeastern fraction of the Turkmeni-
stan part of the Amu Darya Basin (Figure 1) and
investigates the key controls on H
2
S formation and
the occurrence in the greater Charjou terrace area
(Figure 2). The Amu Darya Basin covers an area of
250,000 km
2
(96,525 mi
2
) in the southeastern part of
the Turan platform in central and east Turkmenistan
and western Uzbekistan. It encompasses the Charjou
terrace (or Step), the Bukhara terrace, the Badkhyz
Karabil uplift zone, MaryRepetek uplift zone, Cen-
tral Garagum dome, Bokurdak monocline, Murgab
Depression, Obruchev depression, Beshkent depres-
sion, Karabekaul depression, Zaunguz depression,
andthe Kopet-Dagforedeep(Mashrykov, 1973; Allanov,
1976; Aksenov, 1985; Clarke, 1988; Amursky et al.,
1991). This is a prolific gas and oil province with large
remaining potential. The Charjou terrace is host to
several giant fields, such as the oil-gas-condensate Kok-
dumulak field, with reserves of 5 tcf of gas, 700 MMBO
and 720 million barrels of condensate (MBC) (OCon-
ner and Sonnenberg, 1991), and the gas-condensate
Semanteppe field, with 3.6 tcf gas and 11 MBC (un-
published poster session during Turkmenistan Inter-
national Oil and Gas Exhibition, 1997).
PREVIOUS WORK
Previous work on the gas quality of the Amu Darya
Basin is sparse. After the joint work with Amursky
(1991), Solobyev et al. (1996) provided an overview
of fields with elevated H
2
S contents and areas with
elemental sulfur. They note that H
2
S contents range
upward to 6%, and that great variations are found in
adjacent fields in the same lithofacies zone. For ex-
ample, the H
2
S concentration reaches 3% in the Se-
manteppe and Akkum fields, whereas values of 0.2
0.3% are encountered in the nearby Kishtuvan and
Khodzhikazganfields. Great variations were alsofound
within a single field, e.g., a range from tenths of a
percent to 2.5% in the Kandym field (Uzbekistan).
The authors attribute the variance in H
2
S to elevated
temperature, which has been a favorable factor in
H
2
S formation but do not explicitly attribute this to
a sulfate reduction process. H
2
S-containing gas asso-
ciation with reef traps was also predicted by Am-
maniyazov and Nevmirich (1985).
An evaluation of the hydrocarbon fill and spill
history is presented by Kushnirov et al. (1994). They
134 / Isaksen and Khalylov
attribute the hydrocarbon phase in each trap to the
interdependency of the quantity of the [initial] accu-
mulated oil, the volume of the introduced gas, and
the thermobaric conditions.
DATABASE
The natural gas data discussed herein are mostly
fromthe Institute of Oil and Gas and in part fromthe
State Concern Turkmengeologia Central Laboratory
(Turkmenistan); Geology, Oil and Gas Fields Explo-
rationInstitute, Uzbekistan(IGIRNIGM); UzbekNIPIN-
eftegas (Uzbekistan); and State Enterprise All-Russia
Research Geological Institute of the Ministry of Nat-
ural Resources of the Russian Federation (VNIGNI)
(Moscow). The compositionof hydrocarbonand non-
hydrocarbon gases is shown in Table 1. This includes
compositional data for hydrocarbon and nonhydro-
carbon gases (CO
2
, N
2
, H
2
S, He) and water analyses
(total dissolved solids [TDS] and water densities).
Formation temperatures were from direct measure-
ments in wells and were obtained from both the In-
stitute for Oil and Gas (Turkmenistan) and published
sources (Solobyev et al., 1996).
Analyses include gas chromatography of the nat-
ural gases and standard formation-water tests. Source
rocks were analyzed by Leco total organic carbon
(TOC), Rock-Eval pyrolysis, and for their elemental
composition, whereas solvent extracts were analyzed
by gas chromatography and gas chromatography
mass spectrometry (VNIGNI Labs, Moscow). We ac-
knowledge the lack of more advanced analytical data
(e.g., gas isotopes, x-ray diffraction, and fluid-inclusion
analyses) to investigate detailed chemical reactions
during the thermochemical sulfate reduction (TSR)
process (e.g., Machel et al., 1995a, b; Worden and
Smalley, 1996; Wordenet al., 1995, 1998; Sassen, 1998).
Export of gas samples fromTurkmenistan was not per-
mitted. Nonetheless, this study provides a geological
and geochemical framework, which suggests that the
TSR process may be occurring in the area of study.
GEOLOGICAL SETTING:
SOURCES FOR OIL AND GAS
The Charjou terrace area of the Amu Darya Basin is
host to two main types of source rocks for oil and gas:
Oxfordian calcareous, marine mudstones (the so-
called black-shale clays), and LowerMiddle Jurassic
coaly rocks (Akramhodzhaev, 1977; Akramhodzhaev
and Egamberdiev, 1981; Maksimov et al., 1987). The
Oxfordian calcareous mudstones are present as lami-
nated, ammonite-bearing shales that pinchout against
the larger, isolated reefal buildups like Kokdumulak
and the platform margin. The mudstones accumu-
lated as intermound and distal-marine fines in an
anoxic depositional environment. The development
of anoxic bottom waters is suggested by the presence
FIGURE 1. Map of Turkmenistan showing the location of study area (red oval) in the northeastern part of the Amu Darya
Basin. The I to I
0
line marks the location of the schematic geological cross section shown in Figure 3.
Controls on H
2
S Formation in a Jurassic Carbonate Play, Turkmenistan / 135
of laminated shales with no, or very minor, bioturba-
tion. As a result, organic matter had a greater tenden-
cy to be preserved as oil-prone kerogen as opposed
to being oxidized. Total organic carbon (TOC) con-
tents range upward to 4.5 wt.%, whereas carbonate
(CaCO
3
) contents are in the 4572% range. Hydro-
gen content, a measure of the kerogens ability to gen-
erate oil, was measured by Rock-Eval pyrolysis. The
oil-prone quality of these rocks is given by hydrogen
indices as highas 500mg hydrocarbons/g TOC(Tissot
and Welte, 1984). These values should be considered
as minimum values as temperatures on the Charjou
terrace reach 1201408C in the Oxfordian section.
Therefore, these rocks are within their oil-generative
window and have, as a consequence, experienced
decrease in their hydrogen indices through the pro-
cess of organic maturation. Total sulfur contents in the
rocks range from 1 to 1.4 wt.%. Because SC bonds
(62 kcal/mol) are weaker than CC bonds (82 kcal/
mol) (Pauling, 1970; Powell and Snowdon, 1983; Orr,
1986), these source rocks should start generating oil
at lower temperatures than conventional marine, al-
gal (type II and sulfur-poor) source rocks.
Sterane and triterpane biomarker data from oils in
Uzbekistan suggest that the Oxfordian, organic-rich
marine shales (black shales) are the likely sources
for oils in the fields of the Charjou terrace. Oil-oil
and oil-source correlation studies have been pre-
sented by Maksimov et al. (1987), where the oils in
the southeastern regions of the Bukhara (Uzbeki-
stan) and Charjou (Uzbekistan and Turkmenistan)
areas are interpreted to have been derived from the
marine, sapropelic shales inthe Upper Jurassic. These
oils are characterized as isotopically light (d
13
C26.8
to 29.8%and d
34
S 3.5 to 11.5%), with relatively
high surface densities (0.860.95 g/cm
3
), and en-
riched in sulfur (12%). In contrast, oils from the
northwestern part of the Amu Darya Basin are de-
rived fromLower Middle Jurassic shales witha high
content of terrigenous higher plant organic matter.
These oils are isotopically heavier (d
13
C 26.8 to
25.2% and d
34
S 4.8 to 6.0%), relatively light
FIGURE 2. Location of study area and major gas and gas-condensate fields in the Charjou terrace area of the
northeastern Amu Darya Basin.
136 / Isaksen and Khalylov
Table 1. Hydrocarbon and nonhydrocarbon gas compositions for select fields in the Charjou Terrace, Turkmenistan and Uzbekistan.
Field Well
Number
Base
Depth
(m)
Top
Depth
m)
C
1
(vol.%)
C
2
(vol.%)
C
3
(vol.%)
iC
4
(vol.%)
nC
4
(vol.%)
C
5+
(vol.%)
Gas
Wetness
(%)
CO
2
(vol.%)
N
2
(vol.%)
He
(vol.%)
H
2
S
(vol.%)
Calorific
Value
(kcal/m
3
)
HC
Gas
(%)
Non-HCs
(%)
Akkumulyan 6 2705 2698 95.74 1.5 0.14 0.04 0.035 0.084 1.76 1.13 0.72 0.005 7932 97.539 1.855
Akkumulyan 7 2422 2412 93.59 0.61 0.13 0.02 0.029 0.067 0.84 4.78 0.75 7623 94.446 5.53
Kishtuvan 1 2461 2451 92.75 3.1 0.58 0.09 0.18 3.91 2.67 0.22 96.7 2.89
Akkumulyan 6 2494 2442 92.74 0.87 0.12 0.02 0.03 0.059 1.11 4.16 0.95 1.03 7590 93.839 6.14
Ildjik 4 2228 2200 92.69 3.02 0.63 0.11 0.17 0.25 4.07 1.99 1.12 8133 96.87 3.11
Farab 18 2356 2332 92.66 4.14 0.84 0.16 0.199 0.402 5.45 1 0.58 0.0004 98.401 1.5804
Metedzhen 3 2712 2710 92.5 1.6 0.17 0.12 0.03 2.00 5.4 0.22 94.42 5.62
Yakkui 6 3703 3484 92.41 3.09 0.68 0.14 0.14 0.25 4.20 1.33 1.95 8127 96.71 3.28
Khodzhambaz 5 4093 4081 92.26 3.61 0.49 0.06 0.08 0.04 4.39 2.84 0.62 0.0004 8630 96.54 3.4604
Semanteppe 36 2501 2472 92.19 2.01 0.44 0.08 0.08 0.224 2.75 2.726 1.106 1.127 95.024 4.959
Ildjik 4 2214 2200 91.86 3.34 0.75 0.14 0.22 0.49 4.62 2.32 0.86 8257 96.8 3.18
Akkumulyan 6 2578 2512 91.26 0.94 0.08 0.02 0.02 0.03 1.15 4.096 2.11 0.005 1.4 7531 92.35 7.611
Metedzhen 1 2768 2662 91.1 3.6 0.18 0.08 0.03 4.06 3.9 1.1 0.009 94.99 5.009
Yakkui 2 3845 3490 91.04 3.69 0.81 0.169 0.17 0.34 5.05 1.9 1.84 0.3 8184 96.219 4.04
Bereketli 1 3122 3117 90.679 3.802 0.906 0.179 0.19 0.25 5.30 2.458 1.536 0.023 8159 96.006 4.017
Yankui 2 3472 3468 90.62 3.38 0.73 0.146 0.148 0.207 4.63 2.3 2.458 0.008 8018 95.231 4.766
W. Dzhuramergen 3 3630 3452 90.618 5.11 0.99 0.25 0.28 0.24 6.82 0.21 1.68 0.01 8412 97.488 1.9
Pirgui 4 3300 3294 90.617 3.907 0.889 0.171 0.179 0.216 5.37 2.5 1.43 0.0005 8163 95.979 3.9305
Tangigui 3 3180 3170 90.59 4.25 1.02 0.207 0.221 0.436 5.92 1.984 1.28 0.005 8321 96.724 3.269
Akkumulyan 6 2645 2616 90.32 0.9 0.08 0.02 0.026 0.058 1.12 6.5 2.08 0.005 1.44 7386 91.404 10.025
Tangigui 4 3260 3255 89.71 3.55 0.83 0.154 0.213 0.254 5.03 2.52 2.759 0.0005 8030 94.711 5.2795
Yakkui 1 3644 3631 89.229 3.36 0.78 0.165 0.187 0.28 4.79 4.58 1.026 0.34 7960 94.001 5.946
Bota 1 3337 3331 89 3.2 1.075 0.216 0.319 0.656 5.13 3.104 2.358 0.0025 94.466 5.4645
Metedzhen 1 2759 2637 88.95 1.29 0.16 0.04 0.04 1.65 4.8 0.7 3.8 90.48 9.3
Semanteppe 29 2588 2452 88.697 3.5 1.232 0.198 0.273 0.355 5.54 2.52 2.867 0.01 1.9 8217 94.255 7.297
Bota 1 3232 3224 87.78 3.76 1.026 0.235 0.273 0.853 5.69 4.733 1.33 93.927 6.063
Tangigui 4 3260 3255 87.64 4.8 1.1 0.315 0.369 0.399 6.99 4.2 1.154 0.011 8240 94.623 5.365
Beshir 1 3848 3809 87.15 6 1.89 0.41 0.48 0.56 9.15 1.76 1.74 8653 96.49 3.5
Beshir 1 3844 3784 85.17 5.9 1.77 0.38 0.45 0.74 9.07 3.03 2.56 0.001 8513 94.41 5.591
Girsan 6 3542 3553 85.11 4.9 1.2 0.29 0.3 0.91 7.29 2.38 4.9 8248 92.71 7.28
Ildjik 1 2238 2222 85.06 3.46 1.11 0.23 0.31 0.61 5.67 5.3 0.45 2.94 90.78 8.69
Beshir 1 3795 3792 83.36 5.9 1.86 0.39 0.47 0.62 9.37 3.5 2.4 0.00086 8344 92.6 5.90086
Ildjik 2 2353 2348 82.99 6.3 3.96 0.696 1.12 1.72 12.70 0.63 2.53 0.05 9481 96.786 3.21
GeoMean 4.103279579
Notes: Gas wetness = 100 * (C
2
+ C
3
+ C
4
)/(C
1
+ C
2
+ C
3
+ C
4
).
C
o
n
t
r
o
l
s
o
n
H
2
S
F
o
r
m
a
t
i
o
n
i
n
a
J
u
r
a
s
s
i
c
C
a
r
b
o
n
a
t
e
P
l
a
y
,
T
u
r
k
m
e
n
i
s
t
a
n
/
1
3
7
(0.820.84 g/cm
3
), and have low total sulfur
contents (0.070.5%).
The principal source rocks for gas and
volatile oils in the greater Charjou area are
HettangianBajocian (Lower to Middle Ju-
rassic) and Callovian shales and woody coals
deposited in deltaic to nearshore marine set-
tings. Rocks of this age are relatively thick
and widespread in the eastern Amu Darya
Basin (Figure 3). These rocks are dominated
by terrigenous higher plant organic matter.
These siliciclastic rocks are, by nature, rela-
tively rich in iron, and any free sulfide will
rapidly bind with iron to form pyrite. Con-
sequently, these deltaic to nearshore marine
source rocks should only generate sweet gas.
Today, these sediments are within the gas-
generative window, i.e., at temperatures great-
er than 1508C. Regional geological data, along
withbasinmodeling, suggest that the Charjou
terrace has not undergone uplift, thus infer-
ring that the reservoir and source rock tem-
peratures are at their maximumtemperatures.
FORMATION WATERS
A total of 88 samples of formation waters
from 45 wells were measured for their den-
sities and TDS. Densities range from 1018 to
1236kg/m
3
, witha meanvalue of 1093kg/m
3
,
whereas TDSrangefrom70,000to300,000mg/L,
with a mean value of 123,000 mg/L. Forma-
tion water samples were obtained from car-
bonate rocks interbedded with calcium sul-
fate and halite minerals, as well as postsalt,
Cretaceous, clastic reservoirs (Khodzhakuliev,
1976). Values of TDS about 400,000 mg/L are
common in evaporite-rich basins, such as the
Michigan basin, United States (Hanor, 1994).
Salt-cored structures in the North Sea (e.g.,
Ekofisk) have TDS of 50,000150,000 mg/L
(Egeberg and Aagaard, 1989), whereas the Per-
mian Zechstein carbonate and evaporite se-
quences in the Norwegian North Sea have
TDS values in the same range as those found
in the Charjou terrace (Isaksen, unpublished
data). In our study, the sulfate present in the
formation waters is thought to facilitate the
TSR process. Such sulfate was likely added to
the formation waters from leaching of anhy-
drite, a process that was initiated during, or
shortly after, deposition of these sediments. F
I
G
U
R
E
3
.
S
c
h
e
m
a
t
i
c
g
e
o
l
o
g
i
c
a
l
c
r
o
s
s
s
e
c
t
i
o
n
o
f
t
h
e
e
a
s
t
e
r
n
A
m
u
D
a
r
y
a
B
a
s
i
n
f
r
o
m
S
h
a
t
l
y
k

M
a
r
y
i
n
t
h
e
s
o
u
t
h
a
c
r
o
s
s
t
h
e
C
h
a
r
j
o
u
t
e
r
r
a
c
e
t
o
t
h
e
B
u
k
h
a
r
a
S
t
e
p
i
n
s
o
u
t
h
e
a
s
t
e
r
n
U
z
b
e
k
i
s
t
a
n
.
T
h
e
f
i
g
u
r
e
i
s
m
o
d
i
f
i
e
d
f
r
o
m
A
.
B
a
k
i
r
o
v
(
1
9
7
9
,
p
e
r
s
o
n
a
l
c
o
m
m
u
n
i
c
a
t
i
o
n
)
.
138 / Isaksen and Khalylov
HYDROCARBON MIGRATION
A typical feature of the reservoirs on the Charjou
terrace is the presence of bitumen-staining lining
pores and microfractures (Figure 4). Reservoir tem-
peratures are too low to have caused in-situ thermal
cracking of oil because temperatures greater than
1508C are required for the formation of so-called
pyrobitumens. Instead, the reservoir bitumens are
likely the result of gas de-asphalting and TSR. The
most likely scenario is that these reservoirs were once
charged with black oil gener-
ated from the Oxfordian cal-
careous mudstones. This oil
was subsequentlydisplacedby
gas generated from the later
maturing Lower Middle Ju-
rassic terrigenous source rocks.
As gas enters an oil reservoir,
the high-molecular-weight
compounds become unstable
and settle out of solution as
asphaltenes (bitumens). As an
additional, albeit secondary
process, TSR generates solid
bitumens formed from poly-
merization and/or sulfurized
bitumens (Orr, 1974; Machel,
1987).
Another aspect of hydro-
carbon migration from south
to north across the Charjou
terrace is seen by the grad-
ual increase in condensate
yields and gas wetness. An
increase in condensate yield
occurs from Yangui to Uzun-
guiPirgui to TangiguiBota
to Kokdumulak, which is a re-
gional migration trend with
different pressure and temper-
ature conditions for hydro-
carbons in the presalt Upper
Jurassic (Figure 5). In addi-
tion to the regional increase
in gas wetness from south to
northacross the Charjou Step,
local anomalously high values are present (Figure 5).
These gases have gas-wetness factors (calculated as
100 [C
2+
/C
1
]) between 8 and 16. This indicates that
some of the Oxfordian carbonate reservoirs (Kerven,
Kelyaka, Kokdumalak, Bereketli, and Beshir) are hy-
drostatically isolated, or sheltered, from gas flushing
from the gas-generating HettangianBajocian rocks.
It is likely that separate pressure cells in the Oxfordian
control the subsalt migrationpathways and, thus, help
facilitate the preservation of both wet gases enriched
in volatile oil and oil legs (e.g., Kokdumulak).
FIGURE 4. Thin-section photo-
micrographs of reservoir rocks
showing abundant bitumen
in pore volumes.
Controls on H
2
S Formation in a Jurassic Carbonate Play, Turkmenistan / 139
Figure 6 shows the gas composition compared to
an interpreted level of maturation. As gas isotopic
values were unavailable, evaluations of gas thermal
maturity based on compositional data alone carry a
higher degreeof uncertainty. Maturityevaluations were
made by comparing gas compositions, source rock
facies, and geohistory analyses with analogs from
other basins. Most of the gases are at thermal-maturity
levels equivalent to vitrinite reflectance values of 1.4
1.9% R
o
. This evaluation considers the likely modifi-
cation of gas compositions during TSR (Krouse et al.,
1988), i.e., the added caution that must be applied
FIGURE 6. Gas compositions (volume %) and interpreted levels of thermal maturity. LOM = level of organic maturation.
LOM 9 = 0.65% R
o
; LOM 10 = 0.8% R
o
; LOM 11 = 1.1% R
o
; LOM 12 = 1.5% R
o
; LOM 13.5 = 2% R
o
. C
1
(methane),
C
2
(ethane), C
3
(propane), iC
4
(iso-butane), nC
4
(normal-butane), C
5+
(pentane and higher molecular weight gases).
FIGURE 5. Regional gas wet-
ness values [(C
2
C
5
/C
1
)
100] in the subsalt, Oxfordian
carbonate section in the
Charjou terrace. Legend: 1 =
gas-condensate field; 2 = gas-
condensate-oil fields; 3 = oil
shows; 4 = gas shows; 5 = local
hydrocarbon leads; 6 = gas
wetness values (%) for the
subsalt, CallovianOxfordian
section; 7 = isolines of gas
wetness; 8 = zones with gas
wetness higher than 6;
9 = regional direction of HC
migration in the subsalt
carbonate reservoir; 10 =
Turkmenistan state border.
140 / Isaksen and Khalylov
whenassessinggas-maturitylevels because gas wetness
values decrease as a result of increasing thermal matu-
rity, TSR, and migration-induced fractionation effects.
GAS QUALITY
The composition of hydrocarbon and nonhydro-
carbon gases is shown in Table 1. Hydrocarbon gases
constitute 8998% of the total gas, whereas non-
hydrocarbons constitute 110% (Figure 7). Among
the nonhydrocarbons, CO
2
dominates with second-
ary amounts of N
2
and H
2
S. Helium is reported for
only a few samples and is present in trace amounts.
Methane contents range from 83 to greater than
95%. Gas wetness values range from 1 to 13%, with a
geometric mean at 4%. The gas wetness is calculated
as 100 * ([C
2
+ C
3
+ C
4
]/[C
1
+ C
2
+ C
3
+ C
4
]).
Hydrogen sulfide concentrations range from zero
to 3.8 vol.%. Although not a significant amount of
the total gas volume, these values are nonetheless
high when considering drilling safety and the cost of
facilities required to process the gas to remove H
2
S.
SOURCES OF H
2
S
Four possible sources for H
2
S in the geological
environment exist: TSR, thermal decomposition of
sulfur-rich source rocks, bacterial sulfate reduction
(BSR), and igneous gases (listed inorder of volumetric
importance). H
2
S fromigneous gases is uncommonly
associated with petroleumreservoirs, although deep-
seated faults in certain petroleum provinces may act
as migration conduits. Bacterial sulfate reduction
constitutes enzyme-catalyzed biochemical reactions
that occur at temperatures
below 60808C and is rela-
tively common in recent sed-
iments (Orr, 1977; Machel
et al., 1995a). More specifical-
ly, anaerobic, sulfate-reducing
microorganisms oxidize hy-
drocarbons and organic mat-
ter using sulfate as an oxidant
(Goldstein and Aizenshtat,
1994). These biological reac-
tions cease at higher temper-
atures. It is generally assumed
that the BSR process results
in relatively low amounts of
H
2
S, less than35%(Orr, 1977;
Machel et al., 1995a; Worden
et al., 1995). Factors causing
the amount of H
2
S generated by BSR to remain low
include conversion to elemental sulfur, pyrite, or
organic sulfur (Kaplan et al., 1963; Orr and Gaines,
1974); dissolution in water (Reeburgh, 1969); and the
high toxicity of H
2
S, even to sulfate-reducing mi-
crobes (Sellecket al., 1952). Under very special circum-
stances, such as hot springs and deep-water hydro-
thermal vents, BSR by thermophilic microorganisms
may occur at temperatures as high as 1108C (Orr,
1977; Jrgensen et al., 1992).
In our study area, the 1008C isotherm roughly
parallels the TurkmenistanUzbekistan border at ap-
proximately 25 km (15 mi) north of the border and
makes a southward bend near the Semateppe field. It
follows from the discussion above that the H
2
S pres-
ent in the Uzbekistan fields located north of the
1008C isotherm can be attributed to the BSR process
and/or possible remigration of gas (enriched in H
2
S
by TSR) from the south (Turkmenistan). The latter
process is, however, considered unlikely.
Generationof H
2
S directly fromsource rocks during
thermal maturation is well documented by the de-
tailed studies of the Lacq field in the Aquitaine basin,
France (LeTran, 1974), and the Big Horn basin in
Wyoming (United States) (Orr, 1974). The process
typically involves the generation of H
2
S from ther-
mal degradation of molecular complexes containing
organicallybondedsulfur. Maturationof organic mat-
ter in source rocks, or thermal degradation of crude
oil in high-temperature (>1508C) reservoirs, is, how-
ever, not capable of generating high (>10%) amounts
of H
2
S. Orr (1977) estimated, by mass-balance cal-
culations, that only 23% H
2
S could be generated by
these processes, obviously limited by the amount of
FIGURE 7. Nonhydrocarbon and hydrocarbon gas compositions.
Controls on H
2
S Formation in a Jurassic Carbonate Play, Turkmenistan / 141
organic sulfur in the organic matter undergoing deg-
radation. Furthermore, the concentration of H
2
S will
always decrease during secondary migration because
it reacts to form pyrite or is dissolved in formation
waters. H
2
S will only evolve as a free gas once its con-
centration exceeds its solubility in formation waters.
Within our study area in Turkmenistan, the bulk
volume of gases is generated from Lower to Middle
Jurassic shales andwoody coals depositedindeltaic to
nearshore marine settings. Such type III source rocks
generate sulfur-free gas. Sufficient amounts of reac-
tive iron exist in paralic depositional environments
to bind any free sulfide. Furthermore, H
2
S is lost by
water dissolutionbecause these gases need to migrate
several kilometers from the main gas-producing re-
gions tothe southof the AmuDarya River. In-reservoir
thermal cracking of crude oils trapped on the Char-
jou terrace is unlikely because reservoir (Oxfordian
carbonates) temperatures do not exceed 1508C.
Thermochemical sulfate reduction is considered
the dominant process generating H
2
S in the Charjou
terrace.
THERMOCHEMICAL SULFATE REDUCTION
Thermochemical sulfate reduction is a nonbiologi-
cal process whereby hydrocarbons react with aqueous
sulfate derived from anhydrite (CaSO
4
) to form CO
2
,
H
2
S, and bitumen (Orr, 1974; Goldstein and Aizensh-
tat, 1994; Machel et al., 1995a):
Hydrocarbons SO
2
4
)altered HCs bitumen
HCO

3
CO
2

H
2
S HS
2
heat
Prerequisites for TSR are temperatures greater than
approximately 1008C, the presence of hydrocarbons
or other reactive organic material, and sulfate. The
current understanding of the TSR process is well docu-
mented in the literature. Excellent review papers on
the subject have been published by Trudinger et al.
(1985), Goldstein and Aizenshtat (1994), and Noth
(1997). The lower temperature range for the TSR reac-
tions has been debated in numerous studies (Table 2)
(Machel, 1998; Worden et al., 1998). Case studies of
the PermianTriassic Khuff Formation (Abu Dhabi)
by Worden et al. (1995) showed that only gas reser-
voirs at temperatures above 1408Ccontainlarge quan-
tities (>10%) of H
2
S, and the authors suggested that
previous studies that indicate a lower temperature
limit for the TSR process may have erroneously in-
cluded soured gas that has migrated into structurally
higher positions. However, Orr (1982) found through
laboratory experiments that TSR could occur at lower
temperatures when H
2
S was initially present. Nume-
rous studies on the deep Smackover Formation along
Table 2. Selected case studies reporting on the low-temperature range of TSR reactions.
Reservoir or Field Basin/Area Temperature Range (8C) Reference
McArthur River Area Northern Territory, Australia Approximately 100 Rye and Williams, 1981
Pine Point Pb and
Zn Ore
Northwestern
Territories, Canada
Approximately 100 Macqueen and Powell, 1984
Smackover trend Gulf Coast, U.S.A. 120150 Nunn and Sassen, 1986
Louisiana and
Mississippi
Gulf Coast, U.S.A. 120150 Sassen, 1988
Smackover trend Alabama, Gulf Coast, U.S.A. 130 Claypool and Mancini, 1989
East Smackover Florida, Alabama, Mississippi 100110 Wade et al., 1989
Smackover Formation Southeast Mississippi, U.S.A. 120150 Heydari and Moore, 1989
Devonian Western Canada 100110 Eliuk and Viau, 1992
Nisku Formation Western Canada
sedimentary basin
125145 Machel et al., 1995b
Khuff Formation Abu Dhabi 140 Worden et al., 1995
Wabamun Group/
Leduc Formation
Southwest Alberta, Canada 90175 Krouse et al., 1988
Nisku Formation Brazeau River area, Canada 125145 Manzano et al., 1997
Smackover Formation Mississippi, Gulf Coast, U.S.A. 140150* Heydari, 1997
Charjou terrace Turkmenistan 100110 This study
*Gradual increase from Goodwater field (1158C) to South State Line field (1528C).
142 / Isaksen and Khalylov
the United States Gulf Coast have attributed the high
(as much as 80%) H
2
S concentration to the high-
temperature phase of the TSR process (Wade et al.,
1989; Rhodes, 1994; Heydari, 1997). Studies of the TSR
process in the Western Canada sedimentary basin
have shown that the highest H
2
S concentrations were
associated with very dry gases, implying that ethane
and propane serve as the dominant reducing agents
(Krouse et al., 1988). The high-end temperature range
for TSR is not well constrained, but the process is
knowntooccur toat least 2008C(Machel et al., 1995a).
For a detailed discussion on the reaction path-
ways that may occur during TSR, readers are referred
to Goldstein and Aizenshtat (1994). In short, the key
reactions are
CaSO
4
CH
4
)CaCO
3
H
2
S H
2
O
2CaSO
4
C
2
H
6
)2CaCO
3
H
2
S S 2H
2
O
4S CH
4
2H
2
O )CO
2
4H
2
S
In the Charjou terrace, stylolites seen in cored
sections of the carbonate reservoirs show that mo-
bilization and precipitation of CaCO
3
have occurred.
Several reservoirs have also been altered by dolomi-
tization. Althoughthis is expectedtooccur duringTSR,
it may also occur during normal diagenetic (nonre-
dox) reactions and does not prove that TSR occurred.
A more diagnostic criterion of BSR versus TSR is the
morphology of pyrite. Although pyrite is only present
in trace amounts, its appearance is prismatic instead of
framboidal, thus pointing to the TSR process.
The reactive sulfate for both BSR and TSR reactions
is likely derived from the aqueous dissolution of the
anhydrite and gypsum present in the seal rocks over-
lying most of the carbonate platformand reefal build-
ups. In view of the high (>1008C) temperatures in
these reservoirs, a critical component to TSR is the
lithology of the reservoir seal rock. Traps with anhy-
drite seals have elevated H
2
S contents, whereas
traps capped by the Oxfordian shales and mudstones
have trace levels of, or no, H
2
S, regardless of reservoir
temperature (Figure 8). Fields with elevated H
2
S con-
tents include Semanteppe, Metedzhan, Iljik, and
Akgumolam.
At Iljik, the relation between TSR and the reservoir
andseal lithologies is especially illustrative (Figure 9).
At the Iljik-1 well, H
2
S levels are at 2.9% from test III
at approximately 2047 m(6715 ft) true vertical depth
inthe gas columninclose proximity to the anhydrite
seal. At Iljik-2, test number Vat 19541963 m(6410
6440 ft) true vertical depth, the H
2
S concentration is
zero. This deeper gas column is capped by a high-
radioactive shale and, as a consequence, no TSR has
occurred.
At Akgumolam, the Oxfordian shale serves as the
cap rock facies prestructuring. However, thrusting
has caused the overlying anhydrite to be juxtaposed
to the gas reservoir and, thus, enabling TSR to occur.
H
2
S concentrations are 1.031.4% at Akgumolam-6.
This indicates that large fault surfaces withjuxtaposed
anhydrite can support TSR when other requirements
are favorable.
The Samanteppe field is a broad, simple, dome-
shaped anticline that extends northward. Closure of
the gas-bearing interval is 250 m (820 ft). H
2
S con-
centrations are typically about 24.3% down to the
gas-water contact (2317 m; 7601 ft), which is 250 m
(820 ft) below the lowermost continuous anhydrite
seal. No H
2
S concentration trend exists in the gas
column. The uppermost 100 m(330 ft) of the carbon-
ate reservoir (1.7% porosity, 1.5 md) is interbedded
with thinner anhydrites, followed by bedded (2.7%
porosity and 5 md) and massive (5.7% porosity and
average 100 md) carbonates below. The entire trap is
sealed by anhydrite, including the lateral seals. This
poses the question: How far away from the reactive
site of the anhydrite seal is the effect of TSRobserved?
The H
2
S measurements suggest that the entire gas
column is homogenized, with H
2
S generated by TSR
at the limestone-anhydrite interface at the crest and
flanks of the structure. From regional temperature
gradients, the temperature differential across 250 m
(820 ft) (the height of the gas columnat Semanteppe)
is approximately 4.58C. However, temperature mea-
surements in the field show only a 28C variance in
temperature (specifically 981008C), suggesting that
thermal convection is operative, and that this pro-
cess has reduced the temperature differential ex-
pected over this depth interval. In addition, H
2
S gen-
erated at the gas-anhydrite interface would diffuse
FIGURE 8. Relation of reservoir temperature, H
2
S con-
centrations, and the lithofacies of the seal rock.
Controls on H
2
S Formation in a Jurassic Carbonate Play, Turkmenistan / 143
throughout the reservoir according to Ficks first
law of diffusion. The diffusion coefficient is pro-
portional to the cross-sectional area between the gas
undergoing alteration by TSR and the virgin gas.
Therefore, we infer that the observed homogeni-
zation of H
2
S throughout the main Semanteppe gas
reservoir can be explained by a combination of ther-
mal convection and chemical diffusion.
Among the hydrocarbon gases in the TSR reac-
tions, propane is more reactive than ethane, which
is more reactive than methane (Krouse et al., 1988;
Kiyosuet al., 1990). As a consequence, the TSRprocess
could cause the reservoired gases in close proximity
to anhydrite to become progressively drier (enriched
in methane). This can cause gas wetness to increase
withdepthinthe reservoir, having clear implications
FIGURE 9. Sedimentary section at the Ildjik field in the northwestern part of the Charjou terrace showing geological age,
lithofacies, gamma-ray log, and neutron log. Compartmentalization of the carbonate-dolomite reservoir is likely
controlled by a widespread transgressive-marine black shale (shown here near 2300-m [7500-ft] depth). H
2
S concen-
trations are shown for strata capped by the marine shale and strata capped by anhydrite. The TSR process is inferred to
be operative in the shallow compartment sealed by anhydrite.
144 / Isaksen and Khalylov
for interpretations of gas profiles. An increase of gas
wetness with depth is commonly taken to indicate
proximity to an oil leg, but TSR can cause a similar
effect. Such increases of gas wetness with reservoir
depthare observedat IljikandAkgumolam(Figure 10).
In both of these fields, the gas column is in contact
with anhydrite. The TSR process also generates sig-
nificant amounts of CO
2
. The covariance of H
2
S and
CO
2
, and its dependence on the seal facies, is shown
in Figure 11. This is a good independent indication
that the TSR process is occurring.
As ethane and propane are more reactive than
methane inthe TSR process, a decrease ingas wetness
should covary within increasing concentrations of
both H
2
S and CO
2
(Figure 12). This variance in re-
activity among the hydrocarbon gases is also ex-
pected to influence the temperature range for the
onset of the TSR reactions, i.e., an increase in wetness
should initiate the TSR reactions at lower tempera-
tures. This is only inpart observedwithinthe Charjou
terrace, where an increase in gas wetness above 1.2%
is accompanied by an increase in % H
2
S and a de-
crease in reservoir temperature (Figure 12). Another
variable with poor control is the extent of reservoir
dolomitization. Temperatures in this data set are con-
sidered true maximum temperatures because they
represent equilibrated well shut-in temperatures.
Clearly, a larger data set is required to investigate
this further.
Fromthe findings presented above, it is possible to
predict the gas quality (sour vs. sweet gas) through-
out the Charjou terrace area. The observations illus-
trated in Figure 13 are as follows:
1) Platformal facies in interbedded carbonate and
anhydrite, as well as the larger reefal buildups
sealed by anhydrite, have elevated levels of H
2
S
caused by TSR. Dolomite (CaMgCO
3
) reservoirs
capped by anhydrite have the highest concentra-
tions of H
2
S, indicating that magnesium may act
as a catalyst in the TSR reaction.
2) Structural traps andsmaller reefal buildups draped
by Oxfordian shales have no or trace levels of
H
2
S because the gas column is not in contact
with anhydrite.
3) Some structural traps, such as Akgumolam, are
capped by the marine shales but have anhydrite
juxtaposed to the gas-bearing dolomites because
of thrusting. These traps also have elevated H
2
S
contents.
SUMMARY
1) Hydrocarbon gases trapped in carbonate reser-
voirs of the Charjou terrace area were generated
predominantly from Lower Middle Jurassic
marginal-marine and coastal-plain sources, with
a predominance of terrigenous, higher plant
organic matter. Most of the gases are at thermal-
maturity levels equivalent to vitrinite reflec-
tance values of 1.41.9% R
o
, i.e., within the gas-
generative window of their source rocks. These
gases migrated from their source rocks as sweet
gases.
FIGURE 10. Examples of vertical profiles of hydrocarbon
gas compositions in the Iljik (a) and Akkumulan (b) fields.
FIGURE 11. Relation of the lithofacies of the seal rock and
the concentrations of H
2
S and CO
2
. The TSR process
generates both of these nonhydrocarbon gases.
Controls on H
2
S Formation in a Jurassic Carbonate Play, Turkmenistan / 145
2) H
2
S concentrations are controlled by several
factors, the most important being the lithofa-
cies of the reservoir and seal rocks, temperature,
formation-water chemistry, and gas wetness.
3) H
2
S concentrations are at 14% when an anhy-
drite seal is present. All platform carbonates and
major reefal buildups have elevated H
2
S contents
where dolomite reservoirs are sealed by anhy-
drite. Thermochem-
ical sulfate reduc-
tionappears alsoto
have occurredwith-
in structural traps
where the anhy-
drite has beenstruc-
turally juxtaposed
to the gas-bearing
dolomite reservoir.
4) H
2
S concentrations
are at trace levels
when the reservoir
seal is a shale or
mudstone. In these
cases, the reser-
voired gases are
not in contact with
anhydrite, and the
TSR process is not
operational. The
lack of elevatedcon-
tent of pyrite in the
sealing shales or mud-
stones indicates that these
rocks did not act as a sink
for H
2
S.
5) The combination of
thermal convection and
chemical diffusion has
resulted in a chemical
homogenization of the
larger gas reservoirs and
greatlyreducedanychem-
ical gradient of H
2
S be-
tween the reactive sites
near the anhydrite and
sites at least 150m(492ft)
away from the anhydrite.
6) The TSR process is con-
sidered to be in its ini-
tial stages in the gas res-
ervoirs in the Charjou
terrace.
7) Measured reservoir temperatures are at 941478C
throughout the Turkmenistan part of the Char-
jou Terrace. Our data suggest that the minimum
temperature range for the onset of TSR is 100
1108C, and that the temperature onset of these
reactions appears to be influenced by the gas wet-
ness; i.e., TSR may occur at lower temperatures
when propane, butane, and pentane gases are
FIGURE 12. Relation of reservoir temperature, gas wetness, and H
2
S concentrations
in three fields on the Charjou terrace. As gas wetness increases, the higher reactivity
of propane. butane, and pentane, may have enabled the TSR reactions to occur at
lower reservoir temperatures. Akk = Akkumulan; Sem = Semanteppe.
FIGURE 13. Risk of H
2
S presence in various play types on the Charjou Terrace. The main
controls on H
2
S distribution in these plays appear to be the lithofacies of reservoir and seal
rocks and the reservoir temperature. The greatest risk for H
2
S is in dolomite reservoirs
capped by anhydrite and containing wet gas at temperatures above 1008C.
146 / Isaksen and Khalylov
present. The magnitude of this temperature shift
may be on the order of 10158C as compared to
TSR under dry gas conditions.
ACKNOWLEDGMENTS
We are grateful to the Turkmenistan Cabinet of
Ministers, State Concern Turkmengas, and the Insti-
tute for Oil and Gas in Ashgabat, Turkmenistan, for
permission to publish these results. We also thank
ExxonMobil Exploration Company and study partici-
pants, Mitsubishi and China National Petroleum
Corporation, for permission to release these findings.
The geochemical evaluations of gas quality are
based on data acquired by the Institute for Oil and
Gas, Turkmenistan, and the VNIGNI Institute in
Moscow, Russia.
REFERENCES CITED
Aksenov, A. A., 1985, Amu-Darya oil and gas province, in
Neftegazonostnost podsolevnykh otlozheniy: Mos-
cow, Nedra, p. 6984.
Akramhodzhaev, A. M., 1977, Source rocks of Uzbekistan
and techniques for determining their productivity
potential (in Russian): Moscow, Nedra, 47 p.
Akramhodzhaev, A. M., and M. E. Egamberdiev, 1981,
Main source rocks within Upper Jurassic carbonate
formations of west and south Uzbekistan (in Russian):
Uzbek Geological Journal, v. 5, p. 4756.
Allanov, A. K., 1976, Formation of paleotectonics and oil
and gas prospectivity of Paleozoic and Mesozoic de-
posits of Turkmenistan (in Russian): Moscow, Nedra,
131 p.
Amanniyazov, K. N., and L. E. Nevmirich, 1985, Rift com-
plexes and Late Jurassic deposits of east Turkmenis-
tan and their oil and gas prospectivity (in Russian):
Ashgabat, Turkmenistan, Ilym Press, 224 p.
Amursky, G. I., 1991, A model for H
2
S gas deposits
Examples from Central Asia, in Geology of gas and
gas-condensate accumulations (in Russian): Moscow,
VNIIE Gazprom, 48 p.
Amursky, G. I., K. N. Amanniyazov, and V. M. Ceysler,
1991, Oil and gas prospectivity of Turan and SWGissar
plate boundaries (inRussian): Ashgabat, Turkmenistan,
Ilym Press, 228 p.
Bechtel, A., and W. Puettmann, 1991, The origin of the
Kupferschiefer-type mineralization in the Richelsdorf
Hills, Germany, as deduced from stable isotope and
organic geochemical studies: Chemical Geology, v. 91,
no. 1, p. 118.
Belopolsky, A. V., and M. Talwani, 2007, Assessment of the
Greater Caspian region petroleum reserves and their
role in world energy, in P. O. Yilmaz and G. H. Isaksen,
eds., Oil and gas of the Greater Caspian area: AAPG
Studies in Geology 55, p. 57.
Clarke, J. W., 1988, Petroleum geology of the Amu-Darya
gas-oil province of Soviet Central Asia: U.S. Geological
Survey Open-file Report 88-272, 59 p.
Claypool, G. E., and E. A. Mancini, 1989, Geochemical
relationships of petroleum in Mesozoic reservoirs to
carbonate source rocks of Jurassic Smackover Forma-
tion, southwestern Alabama: AAPG Bulletin, v. 73,
no. 7, p. 904924.
Connan, J., and G. Lacrampe-Couloume, 1993, The origin
of the Lacq Superieur heavy oil accumulation and the
giant Lacq Inferieur gas field, in M. L. Bordenave, ed.,
Applied petroleum geochemistry: Paris, France, Edi-
tions Technip, p. 465488.
Dorian, J. P., 2002, Turkmenistan lists ambitious plans for
gas, oil development: Oil & Gas Journal, October 14,
2002, p. 4346.
Egeberg, P. K., and P. Aagaard, 1989, Origin and evolution
of formation waters from oil fields on the Norwegian
shelf: Applied Geochemistry, v. 4, p. 131142.
Eliuk, L. S., and C. A. Viau, 1992, Porosity in sour gas res-
ervoirs associated with thermochemical sulfate reduc-
tion, in E. George, chair, AAPG 1992 Annual Conven-
tion Proceedings: AAPG and SEPM, abs., 36 p.
Goldstein, T. P., and Z. Aizenshtat, 1994, Thermochemical
sulfate reduction: A review: Journal of Thermal Analy-
sis, v. 42, p. 241290.
Hanor, J. S., 1994, Origin of saline fluids in sedimentary
basins, in J. Parnell, ed., Geofluids: Origin, migration
and evolution of fluids in sedimentary basins: Geolog-
ical Society (London) Special Publication 78, p. 151
174.
Heydari, E., 1997, The role of burial diagenesis in hy-
drocarbon destruction and H
2
S accumulation, Up-
per Jurassic Smackover Formation, Black Creek field,
Mississippi: AAPG Bulletin. v. 81, no. 1, p. 2645.
Heydari, E., and C. H. Moore, 1989, Burial diagenesis and
thermochemical sulfate reduction, Smackover Forma-
tion, southeasternMississippi salt basin: Geology, v. 17,
no. 12, p. 10801084.
Hyne, J. H., 1991, Dictionary of petroleum exploration,
drilling and production: Tulsa, Oklahoma, PennWell
Publishing Company, 625 p.
Jrgensen, B. B., M. F. Isaksen, and H. W. Jannasch, 1992,
Bacterial sulfate reduction above 1008C in deep sea hy-
drothermal vent sediments: Science, v. 258, p. 1756
1757.
Kaplan, I. R., K. O. Emery, and S. C. Ritterberg, 1963, The
distribution and isotopic abundance of sulfur in re-
cent marine sediments off southern California: Geo-
chimica et Cosmochimica Acta, v. 27, p. 297331.
Khodzhakuliev, Y. A., 1976, Hydrogeologic control on for-
mation and location of gas and oil accumulations (in
Russian): Moscow, Nedra, 336 p.
Kiyosu, Y., H. R. Krouse, and C. A. Viau, 1990, Carbon
isotope fractionation during oxidation of light hydro-
carbon gases: Relevance to TSR in gas reservoirs, in
Controls on H
2
S Formation in a Jurassic Carbonate Play, Turkmenistan / 147
W. L. Orr and C. M. White, eds., Geochemistry of sulfur
in fossil fuels: American Chemical Society Symposium
Series 429, p. 633641.
Krouse, H., C. A. Viau, L. S. Eliuk, A. Ueda, andS. Halas, 1988,
Chemical and isotopic evidence of thermochemical
sulfate reduction by light hydrocarbon gases in deep
carbonate reservoirs: Nature, v. 333, no. 2. p. 415
419.
Kushnirov, V. V., T. I. Ubaykhodzhayev, and V. I. Sokolov,
1994, Use of various parameters of hydrocarbon gas-
liquid systems for optimization of oil-gas exploration
in carbonate productive complexes: Geologiya Nefti i
Gaza, v. 9, p. 1317.
LeTran, K., 1974, Diagenesis of organic matter and occur-
rence of hydrocarbons and hydrogen sulfide in the
SW Aquitaine basin (France): Bulletin du Centre de
Recherches Pau, v. 8, p. 111137.
Machel, H. G., 1987, Some aspects of diagenetic sulphate-
hydrocarbon redox reactions, in J. D. Marshall, ed.,
Diagenesis of sedimentary sequences: Geological So-
ciety (London) Special Publication 36, p. 1528.
Machel, H. G., 1998, Gas souring by thermochemical sul-
fate reduction at 1408C: Discussion: AAPG Bulletin,
v. 82, no. 10, p. 18701873.
Machel, H. G., H. R. Krouse, and R. Sassen, 1995a, Products
and distinguishing criteria of bacterial and thermo-
chemical sulfate reduction: Applied Geochemistry,
v. 10, p. 373389.
Machel, H. G., H. R. Krouse, L. R. Riciputi, and D. R. Cole,
1995b, Devonian Nisku sour gas play, Canada: A
unique natural laboratory for study of thermochem-
ical sulfate reduction, in M. A. Vairavamurthy and
M. A. A. Schoonen, eds., Geochemical transformations
of sedimentary sulfur: American Chemical Society
Symposium Series, v. 612, p. 439454.
Macqueen, R. W., and T. G. Powell, 1984, Genesis of Pine
Point lead-zinc ores by organic matterSulphate reac-
tions, in Geological Association of Canada and Miner-
alogical Association of Canada, Joint Annual Meeting,
v. 9, p. 86.
Maksimov, S. P., R. G. Pankina, and A. M. Smakhtina,
1987, Conditions of formation of hydrocarbon ac-
cumulations in Mesozoic sediments of the Amudarya
gas-oil province: Geologiya Nefti i Gaza, v. 5, p. 20
27.
Manzano, B. K., M. G. Fowler, and H. G. Machel, 1997,
The influence of thermochemical sulfate reduction on
hydrocarbon composition in Nisku reservoirs, Bra-
zeau River area, Alberta, Canada: Organic Geochem-
istry, v. 27, no. 78, p. 507521.
Mashrykov, K. K., 1973, Geology of Turkmenistan (in Rus-
sian): Ashgabat, Turkmenistan, Ilyum Press, 198 p.
Noth, S., 1997, High H
2
S contents and other effects of
thermochemical sulfate reduction in deeply buried car-
bonate reservoires: A review: Geologische Rundschau,
v. 86, p. 275287.
Nunn, J. A., and R. Sassen, 1986, The framework of hydro-
carbon generation and migration, Gulf of Mexico con-
tinental slope: Transactions of the Gulf Coast Associ-
ation of Geological Societies, v. 36, p. 257262.
OConner Jr., R. B., and S. Sonnenberg, 1991, Amu-Daria
liquids potential indicated: Oil & Gas Journal, v. 89,
no. 22, p. 104110.
Orr, W. L., 1974, Changes in sulfur content and isotopic
ratios of sulfur during petroleum maturation study of
Big Horn basin Paleozoic oils: AAPG Bulletin, v. 58,
no. 11, p. 22952318.
Orr, W. L., 1977, Geologic and geochemical controls onthe
distribution of hydrogen sulfide in natural gas, in
R. Campos and J. Goni, eds., Advances in Organic Geo-
chemistry, 1975: Madrid, Spain, Enadimsa, p. 571597.
Orr, W. L., 1982, Rate and mechanism of non-microbial
sulfate reduction: 95th Geological Society of America
Annual Meeting Abstracts Paper 213, p. 580.
Orr, W. L., 1986, Kerogen/asphaltene/sulfur relationships
in sulfur-rich Monterey oils, in D. Leythaeuser and
J. Rullkotter, eds., Advances in organic geochemistry,
1985: Organic Geochemistry, v. 10, p. 499516.
Orr, W. L., and A. G. Gaines Jr., 1974, Observations on rate
of sulfate reduction and organic matter oxidation in
the bottom waters of an estuarine basin: The upper
basin of the Pettaquamscutt River (Rhode Island):
Abstracts of ReportsInternational Congress on Or-
ganic Geochemistry: 6. Advances in organic geochem-
istry 1973, p. 791812.
Pauling, L., 1970, General chemistry, 3d ed.: San Francisco,
California, Freeman, 913 p.
Powell, T. G., and L. R. Snowdon, 1983, A composite hy-
drocarbon generation model: Implications for evalu-
ation of basins for oil and gas: Erdol und Kohle-Erdgas-
Petrochemie vereinigt mit Brennstoff-Chemie, v. 36,
p. 163170.
Reeburgh, W. S., 1969, Observations of gases in Chesa-
peake Bay sediments: Limnology and Oceanography,
v. 14, p. 368375.
Rhodes, J. A., 1994, Reservoir property changes caused by
thermochemical sulfate reduction in the Smackover
Formation: Transactions of the Gulf Coast Association
of Geological Societies, v. 44, p. 605610.
Rye, D. M., and N. Williams, 1981, Studies of base metal
sulfide deposits at McArthur River, Northern Territo-
ry, Australia: III. The stable isotope geochemistry of
the H.Y.C. Ridge and Cooley deposits: Economic Ge-
ology and the Bulletin of the Society of Economic
Geologists, v. 76, no. 1, p. 126.
Sassen, R., 1988, Geochemical and carbon isotope studies
of crude oil destruction, bitumen precipitation, and
sulfate reduction in the deep Smackover Formation:
Organic Geochemistry, v. 12, p. 351361.
Selleck, F. T., L. T. Carmichael, and B. H. Sage, 1952, Phase
behavior of the hydrogen sulfide-water system: Indus-
trial and Engineering Chemistry, v. 44, p. 22192226.
Solobyev, N. N., V. A. Kuzminov, and L. S. Salina, 1996,
Prospects for exploration for gas accumulations in
southern regions of Turan platform: Geologiya Nefti i
Gaza, v. 9, p. 1723.
148 / Isaksen and Khalylov
Tissot, B., and D. Welte, 1984, Petroleum formation and
occurrence: Springer Verlag, Berlin, 699 p.
Trudinger, P. A., L. A. Chambers, and J. W. Smith, 1985,
Low-temperature sulfate reduction: Biological versus
abiological: Canadian Journal of Earth Sciences, v. 22,
p. 19101918.
Wade, W. J., J. S. Hanor, and R. Sasses, 1989, Controls on
H
2
S concentration and hydrocarbon destruction in the
eastern Smackover trend: Transactions of the Gulf Coast
Association of Geological Societies, v. 39, p. 309320.
Worden, R. H., and P. C. Smalley, 1996, H
2
S-producing re-
actions in deep carbonate gas reservoirs: Khuff For-
mation, Abu Dhabi: Chemical Geology, v. 133, p. 157
171.
Worden, R. H., P. C. Smalley, and N. H. Oxtoby, 1995, Gas
souring by thermochemical sulfate reduction at 1408C:
AAPG Bulletin, v. 79, no. 6, p. 854863.
Worden, R. H., P. C. Smalley, and N. H. Oxtoby, 1998, Gas
souring by thermochemical sulfate reductionat 1408C:
Reply: AAPG Bulletin, v. 82, no. 10, p. 18741875.
Controls on H
2
S Formation in a Jurassic Carbonate Play, Turkmenistan / 149

You might also like