Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Marine and Petroleum Geology 78 (2016) 76e87

Contents lists available at ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Research paper

Geochemical characteristics and isotopic reversal of natural gases in


eastern Kopeh-Dagh, NE Iran
Hossein Saadati a, *, Hayder J. Al-Iessa a, Bahram Alizadeh a, b, Elham Tarhandeh c,
Mohammad Hasan Jazayeri c, Hossein Bahrami c, Mehrab Rashidi c
a
Department of Geology, Faculty of Earth Sciences, Shahid Chamran University of Ahvaz, Iran
b
Petroleum Geology and Geochemistry Research Centre (PGGRC), Shahid Chamran University of Ahvaz, Iran
c
National Iranian Oil Company, Exploration Directorate, Tehran, Iran

a r t i c l e i n f o a b s t r a c t

Article history: Natural gas samples from two gas fields located in Eastern Kopeh-Dagh area were analyzed for mo-
Received 19 May 2016 lecular and stable isotope compositions. The gaseous hydrocarbons in both Lower Cretaceous clastic
Received in revised form reservoir and Upper Jurassic carbonate reservoir are coal-type gases mainly derived from type III
17 August 2016
kerogen, however enriched dD values of methane implies presence of type II kerogen related material
Accepted 5 September 2016
Available online 7 September 2016
in the source rock. In comparison Upper Jurassic carbonate reservoir gases show higher dryness co-
efficient resulted through TSR, while presence of C1eC5 gases in Lower Cretaceous clastic reservoir
exhibit no TSR phenomenon. Carbon isotopic values indicate gas to gas cracking and TSR occurrence in
Keywords:
Carbon isotope
the Upper Jurassic carbonate reservoir, as the result of elevated temperature experienced, prior to the
Eastern Kopeh-Dagh following uplifts in last 33e37 million years. The d13C of carbon dioxide and d34S of hydrogen sulfide in
TSR Upper Jurassic carbonate reservoir do not primarily reflect TSR, as uplift related carbonate rock
Mixing dissolution by acidic gases and reaction/precipitation of light H2S have changed these values severely.
Gaseous hydrocarbons in both reservoirs exhibit enrichment in C2 gas member, with the carbonate
reservoir having higher values resulted through mixing with highly-mature-completely-reversed shale
gases. It is likely that the uplifts have lifted off the pressure on shale gases, therefore facilitated the
migration of the gases into overlying horizons. However it appears that the released gases during the
first major uplift (33e37 million years ago) have migrated to both reservoirs, while the second
migrated gases have only mixed with Upper Jurassic carbonate reservoir gases. The studied data
suggesting that economic accumulations of natural gas/shale gases deeper than Upper Jurassic car-
bonate reservoir would be unlikely.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction Kopeh-Dagh foredeep, and near the boundary with the Badkhyz-
Maimana uplift.
Southwest part of the Amu-Darya Basin is locate in NE Iran Lower to Middle Jurassic coaly clastics and coals, with vitrinite
which is called as Eastern Kopeh-Dagh (Fig. 1). This basin is reflectance ranging 1.15Ro percent at the top (3000 m) to
bounded by the Kopeh-Dagh foldbelt along with the frontal fault of 2.30e2.40Ro percent at depths of 4600e5500 m (Kuleshov and
the Kopeh-Dagh (Ulmishek, 2004). The Amu-Darya Basin is ranked Ignatova, 1990; Ulmishek, 2004), along Upper Jurassic (primarily
15th among 102 provinces for its original reserves (USGS, 2000). Oxfordian) basinal black shales have known as source rocks in the
The largest gas reserves are located in the Murgab depression and Amu-Darya Basin (Brookfield and Hashmat, 2001; Ulmishek,
northwest slope of the Badkhyz-Maimana uplift (Fig. 1) (Ulmishek, 2004), however the coal beds of the Upper Triassic sequence in
2004). The Iranian discovered reserves are found mostly in the the Agh-Darband could also be considered as the source rock
(Zanchetta et al., 2013). Due to early diagenesis, the Kashafrud
Formation is do not qualified as reservoir but is considered as one
of the main source rocks of the Amu Darya Basin (Poursoltani and
* Corresponding author. þ989144565399.
Gibling, 2011). Whole thickness of this Formation have not been
E-mail address: hsaadati29@gmail.com (H. Saadati).

http://dx.doi.org/10.1016/j.marpetgeo.2016.09.004
0264-8172/© 2016 Elsevier Ltd. All rights reserved.
H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87 77

Fig. 1. General geological map of the studied area, NE Iran and it situation to Amo Darya basin after (Ulmishek, 2004).

drilled but some intervals are considered as a potential shale between oils from fields in different stratigraphic intervals than
(unpublished data). Nevertheless, the organic-rich calcareous those same stratigraphic interval with different areas (Ulmishek,
shales (marlstone) and micritic limestone of the Middle Jurassic 2004). Thermogenic gases in Cretaceous rocks are drier and
Chaman-Bid Formation with total organic carbon (TOC) contents contain lower contents of heavier homologues with little or no
of about 1.1e2.6% is introduced as the source of the gas in the sulfur (Semenovich et al., 1983) in comparison to that of Upper
eastern part of the Kopeh-Dagh basin (Afshar-Harb, 1979) (Fig. 2). Jurassic carbonates, which have variable amount of condensate and
Most Importantly, the Chaman-Bid Formation have not reached commonly more sulfur (Ulmishek, 2004). The lithofacies of the
proper thermal maturity until early Tertiary (late Paleocene: reservoir and seal rocks along reservoir temperature control the
Moussavi-Harami and Brenner, 1992). distribution of the hydrogen sulfide (H2S; Isaksen and Khalylov,
Biomarkers data from condensates show high level of thermal 2007).
maturity (gas-window zone) and terrestrial organic matter for The gas components and their isotope compositions provide a
source rocks (Sokolova et al., 1993). There are more similarity comprehensive reflection of the sources/paleoenvironments,
78 H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87

Fig. 2. New stratigraphic chart of the Kopeh Dagh belt and of two main adjacent Amu Darya and South Caspian Sea basins (Robert et al., 2014).

organic matter type, evolution degree and processes such as their origin and some secondary processes such as mixing and
migration and the isotopic fractionation effects made happened by thermochemical sulfate reduction (TSR).
the biological, physical and chemical processes through the
geological time. Obtaining comprehensive information about
2. Geological setting
presence or absence of hydrocarbon in an area through geochem-
ical studies would reduce the economic risks majorly. This study
The Kopeh-Dagh is an intracontinental basin which formed after
investigates gaseous hydrocarbon from two structures (A and B) in
Paleo-Tethys closure during Triassic-Jurassic boundary (Golonka,
the Eastern Kopeh-Dagh (Kopeh-Dagh foredeep; Fig. 1) to clarify
2004; Fürsich et al., 2009). Sedimentation started with deposition
H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87 79

of the Middle Jurassic Kashafrud Formation in the rifting basin. and after flushing the lines for 15e20 min to remove air. The
From Middle Jurassic to Miocene beside five major transgressive- chemical and isotopic compositions of gaseous hydrocarbon were
regressive sequences (Moussavi-Harami and Brenner, 1992) over determined on a Finnigan Mat Delta S mass spectrometer inter-
6 km of sedimentary rocks were deposited in the basin (Afshar faced with an HP 5890II gas chromatograph. Gas components were
Harb, 1979). Absence of the Kashafrud Formation from preexist- separated on the gas chromatograph in a stream of helium, con-
ing basement highs and thousands meters thick on hanging walls verted into CO2, and then injected into the mass spectrometer. The
are due to a rifting system (Robert et al., 2014). Presence of stable carbon isotope data are expressed in the delta notation in
erosional surface on top of the Middle Jurassic Kashafrud Formation permil (‰) relative to VPDB (Vienna Pee Dee Belemnite,
indicates brief intra Jurassic hiatus which is followed by the d13CVPDB ¼ 0‰). Measurement precision is ±0.5‰ for d13C. The
deposition of the main reservoirs e.g. Upper Jurassic carbonate stable hydrogen isotope data are reported in delta notation with
reservoir and Lower Cretaceous clastic reservoir formations (Fig. 2) respect to VSMOW (Vienna Standard Mean Ocean Water,
(Robert et al., 2014). Continuous sedimentation was established dDVSMOW ¼ 0‰). Analytical precision is estimated to be ±3‰.
until upper Cretaceous where Turonian and Maastrichtian is
distinguished by general hiatus in the whole Kopeh-Dagh region. 4. Results
The main inversion phase recorder in the Kopeh-Dagh occurred
during the Late Eocene (Frizon de Lamotte et al., 2011) and resulted 4.1. Gas geochemistry
in folding of marine sediments up to the Late Eocene during the
Alpine phase (Robert et al., 2014). Finally, the tectonic rearrange- Geochemical parameters of gas samples from two gas fields are
ment in North Iran started during the Pliocene age led to privilege listed in Table 1. The gases from Lower Cretaceous clastic reservoir
of the strike-slip component in the present-day deformation active are mainly composed of hydrocarbon gaseous, with non-
of Kopeh-Dagh. hydrocarbons content <5% (H2S, N2 and CO2). The CH4 and wet
Upper Jurassic carbonate succession is considered as the main gases range from 93.78 to 94.6% and 2.2e2.6%, respectively. The
reservoir in the Eastern Kopeh-Dagh. It is composed mainly of gases are dry with dryness coefficients (C1/C1eC5) of 0.971e0.976
limestones and dolomites. However, marl/shales, sandstones and (Table 1). The H2S content in Lower Cretaceous clastic reservoir is
evaporates occur to a lesser extent (Afshar Harb, 1979). It is proved negligible (0.05) while N2 and CO2 contents are 2.34e2.74% and
that the main hydrocarbon pay zones had an aragonite to high-Mg <1%, respectively. The natural gases of Upper Jurassic carbonate
calcite mineralogy deposited during the Kimmeridgian (Kavoosi, reservoir display higher dryness coefficient (C1/C1eC5) of >0.99,
2014). Hence, reservoir quality is controlled by depositional facies contributed by extremely lower content of wet gases (Table 1).
and diagenetic processes such as dolomitization and dissolution. These contents as well as non-hydrocarbons content are generally
Increase in the dolomitization and intercrystalline porosity would higher than those of gases from clastic reservoir with CO2 and H2S
appear to be occurred at the late of highstand systems tract (HST) contents in the range of 5.3e5.9% and 2.1e2.6%, respectively, the N2
particularly where anhydrites cap the parasequences (Kavoosi contents are lower within the range of 1.2e1.5% (Table 1).
et al., 2009; Kavoosi, 2014). It seems that this dolomite type
formed at high temperatures, ranging from 94 to ~ 110  C (Adabi,
4.2. Carbon and hydrogen isotopes of gaseous alkanes
2009). Secondary gas reservoir in the Kopeh-Dagh Basin is Lower
Cretaceous clastic sequences with siliciclastic lithology and sub-
Isotopic compositions of natural gases from clastic reservoir are
litharenitic redbeds in composition (Moussavi-Harami and
different from that of carbonate reservoir. Collected gases from
Brenner, 1993; Mortazavi et al., 2014). Some process such as early
clastic reservoir contained sufficient quantity of C1-5 gases for iso-
clay infiltration, and extending silica, calcite and anhydrite
topic analyses; while the isotopic values were measured only on C1-
cementation have occurred in the Lower Cretaceous clastic reser-
3 n-alkanes in Upper Jurassic carbonate reservoir (Table 2). The
voir, resulting in poor reservoir quality. Furthermore clay-sized
d13C1, d13C2, d13C3, d13C4 and d13C5 values of natural gases in the
hematite particles coat most of the grains (Moussavi-Harami and
Lower Cretaceous clastic reservoir range from 34.4 to 31.8‰,
Brenner, 1993). Diagenetic events led to partial dissolution of
from 21.2 to 18.8‰, from 22.9 to 20‰, from 22 to 19.1‰
early and late-stage calcite cements that created secondary
and from 22.1 to 17.2‰, respectively (Table 2), whereas these
porosity in the sandstone units.
values for C1, C2 and C3 of natural gases in Upper Jurassic carbonate
reservoir are from 35.7 to 32.1‰, from 15.4 to 13.6‰ and
3. Samples and methods from 21.8 to 21.3‰, respectively (Table 2). A partial reversal of
carbon isotope values of gaseous alkanes is recognizable in both gas
All samples collected from production wells using steel bottles reservoirs, with Upper Jurassic reservoir gases exhibiting greater

Table 1
Chemical composition of gases.

Field and reservoir CH4 C2H6 C3H8 iC4H10 nC4H10 nC5H12 C2H6/C3H8 Wet Gas C1/C2þC3 N2 H2S CO2 DI% CHC Wetness

B-Clastic-B 93.82 1.41 0.42 0.17 0.25 0.27 3.34 2.34 51.13 2.74 0.05 0.71 97.37 51.13 2.34
B-Clastic-B 94.33 1.42 0.43 0.18 0.26 0.26 3.26 2.39 50.82 2.33 0.05 0.61 97.35 50.82 2.39
B-Clastic-D 93.99 1.71 0.40 0.11 0.18 0.20 4.25 2.71 44.29 2.63 0.05 0.61 97.29 44.29 2.71
B-Clastic-D 93.78 1.69 0.45 0.13 0.22 0.24 3.70 2.62 43.54 2.69 0.05 0.6 97.14 43.54 2.62
A-Clastic-D 94.66 1.59 0.33 0.08 0.12 0.12 4.72 2.22 49.00 2.39 0.05 0.6 97.66 49.00 2.22
A-Clastic-D 93.93 1.69 0.42 0.12 0.20 0.23 3.96 2.56 44.16 2.4 0.05 0.8 97.21 44.16 2.56
B-Carbonate 89.78 0.56 0.06 0.01 0.03 0.03 8.69 0.75 142.52 1.51 2.62 5.35 99.29 142.52 0.75
B-Carbonate 90.15 0.56 0.06 0.01 0.03 0.03 8.72 0.75 142.64 1.22 2.13 5.77 99.22 142.64 0.75
B-Carbonate 89.96 0.56 0.06 0.01 0.03 0.03 8.71 0.75 142.58 1.21 2.43 5.66 99.22 142.58 0.75
B-Carbonate 89.98 0.55 0.06 0.01 0.031 0.03 8.55 0.74 144.91 1.42 2.11 5.56 99.23 144.91 0.74
B-Carbonate 89.67 0.57 0.06 0.01 0.031 0.03 8.64 0.76 141.00 1.32 2.45 5.91 99.21 141.00 0.76

Dryness Indies¼(C1/(C1eC5)*100; CHC ¼CH4/(C2H6þC3H8).


80 H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87

Table 2
Stable isotopic compositions of natural gas from Eastern Kopeh-Dagh.

Field and reservoir d13 d13 a 13


d d13 d13 d13 dD d13 d34S
CH4 C2H6 C2H6 C3H8 nC4H10 nC5H12 CH4 CO2 H2S

B- Clastic -B 31.8 19.6 24.6 22.4 20.9 17.2 163 6.5 e


B- Clastic -B 34.4 20.6 25.6 22.9 20.2 19.3 158 6.6 e
B- Clastic -D 34.2 20 25 21.2 20.1 17.7 163 7.2 e
B- Clastic -D 33 19 24 20.4 19.5 18.7 168 7.8 e
A- Clastic -D 32.9 18.8 23.8 20 19.1 19.4 165 7 e
A- Clastic -D 33.8 21.2 26.2 22.6 22 22.1 161 7.3 e
B- Carbonate 32.7 14.1 e 21.3 e e 158 0 16.5
B- Carbonate 35.7 15.4 e 21.7 e e 163 0.4 16.7
B- Carbonate 33.9 14.1 e 21.5 e e 163 1 16.2
B- Carbonate 32.1 13.6 e 21.8 e e 163 1.1 17
B- Carbonate 32.5 14.1 e 21.5 e e 163 2.1 16.8
a
Initial d13C value of C2H6 plus lower end (5‰) of ethane heavier values than methane ¼ (d13C2þ5).

reversal. The dD1 value of Lower Cretaceous clastic reservoir gases display isotopic values heavier than 28‰ and 25‰ respectively
ranges from 168 to 158‰ (average 163‰), while they are (enriched in d13C; Stahl and Carey, 1975; Dai et al., 1992, 2005).
from 163 to 158‰ (average 162‰) in gases analyzed from Isotopic values of ethane in all analyzed samples as well as the
Upper Jurassic carbonate reservoir (Table 2). location of the plotted data on Fig. 3, characterize the natural
The d13CCO2 and d34SH2S values for Upper Jurassic carbonate gases from both reservoirs as thermogenic gases from humic
reservoir vary from 0.0 to 2.1‰ (average 0.6) and from 16.2 to 17‰ source (higher values for ethane in Figs. 3 and 5 are due to
respectively (average 16.6‰; Table 2). The d13CCO2 values for Lower migration of gases with reversal properties into both of the res-
Cretaceous clastic reservoir range from 6.5 to 7.8‰ ervoirs, which will be discussed later) and confirmed by higher
(average 7.06‰), however the d34SH2S values were not detectable values of d13C3 (Table 2). Generally ethane in relation to methane
due to low amounts of H2S gases. is heavier by 5‰e10‰ PDB (Silverman, 1971; Deines, 1980).
Considering the lower end (d13Ce5, and maintaining the ther-
mogenic gases trend d13C1 < d13C2 < d13C3) isotopic values of
5. Discussion
ethane should have been around d13C 26 or 27‰ prior to
migration/mixing. These calculated ethane values put the gases in
5.1. Genetic types of gaseous alkanes
the area of coal type gas (Table 2), the difference between present
and calculated/original values of ethane implies mixing with
Based on the chemical compositions and isotopic values of
highly mature gases and/or result of thermochemical sulfate
methane and its homologues the type (oil-type or coal-type) of
reduction (TSR). Furthermore during humic kerogen maturation
natural gas could be recognized (Stahl and Carey, 1975; Galimov,
history, higher amounts of thermogenic methane are generated in
1988; Yongchang and Ping, 1996; Dai et al., 2005; Liu et al.,
comparison to C2þ components, whereas sapropelic kerogens
2008). Generally thermogenic gases generated from sapropelic
generate significant amounts of C2þ hydrocarbons (Hunt, 1996),
source rocks are characterized by d13C2 and d13C3 values lighter
which strengthens the idea that the gases in both reservoirs are
than 28‰ and 25‰, respectively (depleted in d13C), whereas
humic in nature (Table 1).
ethane and propane gases derived from humic source rocks

Fig. 3. d13C1 versus d13C2 of the natural gases. Data plot introduce the gases as coal-type gas, enriched ethane gas values of Upper Jurassic reservoir gases are due to mixing with
gases contained highly enriched-ethane.
H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87 81

Gas dryness index is commonly used by researchers to express (Fig. 5). All gases from Lower Cretaceous clastic reservoir fall in the
the volume of methane gases versus other n-alkane homologues. area of primary cracking gas (excluding the effects of mixed gases
Over the years researchers have used different methods of calcu- with higher ethane isotopic values, which will be discussed later)
lating gas dryness; C1/SC1-C5 (Tissot and Welte, 2012), C1/SC2-C5 indicating these gases as primary that have not experienced
(Cody and Hutcheon, 1994; Draper and Boreham, 2006) and C1/ enough temperature to start cracking into secondary gases. On the
(C2 þ C3) (Bernard et al., 1976, 1978; Hunt, 1996; Smith and other hand due to higher maturation, mixing and/or TSR, the
Pallasser, 1996; Whiticar, 1996, 1999; Faiz and Hendry, 2006; d13C2e d13C3 values of Upper Jurassic carbonate reservoir natural
Flores et al., 2008; Osborn and McIntosh, 2010; Stra˛ po c et al., gases are un-plot-able on the diagram. (The reasons for this
2008, 2011). Calculated dryness coefficients are presented in observation will be discussed later), however C2H6/C3H8 values of
Table 1. It is clearly evident that the chemical and carbon isotope Upper Jurassic carbonate reservoir gases are comparable with the
compositions of the natural gases in the Lower Cretaceous clastic C-III area, implying that Upper Jurassic carbonate reservoir gases
reservoir are different from those of Upper Jurassic carbonate are secondary thermal cracking products of previous gases (Fig 5).
reservoir. Higher dryness and heavier carbon isotope values of It is believed that the hydrogen isotopic value of the methane is
gaseous n-alkanes (butane and pentane content were below the related to the 1) depositional environment of the source rock and 2)
detection limit) address Upper Jurassic carbonate reservoir gases as maturity (Dai et al., 1992; Guoyi et al., 2010). This value like carbon
highly mature dry-gas (Hao et al., 2008; Ma et al., 2008; Liu et al., isotope generally increases (becoming less negative) with
2013). increasing carbon number of the gases. The hydrogen isotopic value
The ratio of methane to the sum of ethane and propane (C1/ of thermogenic methane generated from marine source rocks is
(C2 þ C3)) versus d13CeCH4 of natural gas is widely used in dis- higher than 180‰, on the other hand gases generated from
tinguishing the origin (microbial, thermogenic or a mixture of the terrigenous, freshwater environment source rocks have values
two ends) and the type (Kerogen type II or III) of the natural gas lower than 190‰, while gases from paralic, brackish water
(Bernard et al., 1978; Osborn and McIntosh, 2010; Stra˛ poc et al., environment values rank in between (Schoell, 1980). Hydrogen
2008, 2011). A modified Bernard diagram characterizing natural isotope values of gases from Lower Cretaceous clastic reservoir
gases from both gas reservoirs is shown in Fig. 4. Samples collected range from 168 to 158‰ (average 163‰), while Upper Jurassic
from clastic reservoir were plotted in the area for gases from carbonate reservoir gases have values from 163 to e 158‰
kerogen type III, in contrast gases from Upper Jurassic carbonate (average 162‰ Table 2). These values from both reservoirs are
reservoir were plotted close to the area of kerogen type II. Even comparable to marine source rocks, despite previous results indi-
though Figs. 3 and 5 show these gases to be derived from kerogen cating terrigenous coal-derived gases and the fact that the majority
type III with higher maturity. The data does not follow the of the source rocks in the studied area were considered to be
increasing thermal maturity trend, the reasons for this observation kerogen type III source rocks (Ulmishek, 2004). Therefore possible
could be related to the effects of alteration by TSR (Hao et al., 2008) explanations could be that the gases have generated from a source
and also mixing of gases with different maturity. rock consisted mainly from terrigenous organic matter along with
With increasing temperature, primary kerogen and other hy- marine organic matter, or terrigenous source rock co-existing with
drocarbons (including oil, condensate or even wet gases) will start transitional facies/source rocks deposited through previous trans-
cracking into secondarily products. In order to identify whether the gression/regression sequences (Chen, 1994; Dai et al., 2005; Liu
gases are primary or secondary in origin, a modified Prinzhofer et al., 2015). While the isotopic values of the gases from two res-
model (Prinzhofer and Huc, 1995) of C2/C3 and d13C2ed13C3 is used ervoirs show great differences in the carbon isotope values, the
hydrogen isotopic values are almost similar (Table 2). As it was
mentioned earlier the hydrogen values are controlled by deposi-
tional environment and thermal maturity. This observation implies
that the mechanism responsible for the observed differences in
Upper Jurassic carbonate reservoir is secondary processes such as
TSR and mixing rather than thermal cracking.

5.2. In-reservoir mixing of gaseous hydrocarbons

Chung et al. (1988) suggested that gas components generated


from a single source exhibit a semi-linear isotope trends when
stable carbon isotopic compositions of methane and wet gas
(C2eC5) components are plotted against their reciprocal carbon
number (known as Chung plot and/or natural gas plot). As shown in
Fig. 6, the carbon isotopic series of alkanes for both reservoirs
exhibit partial reversals. The natural gases from Lower Cretaceous
clastic reservoir contain complete set of n-alkanes (C1eC5), while
propane is the heaviest n-alkane in Upper Jurassic carbonate
reservoir to be plotted. Both reservoirs show partial reversal in
ethane isotopic values (isotopic reversal is greater in Upper Jurassic
carbonate reservoir; Fig. 6), while isotopic values of methane and
propane in two reservoirs are similar, implying that the source of
gases in both reservoirs is likely to be the same. Geologically clastic
reservoir is located above carbonate reservoir which strengthen the
idea that the gases have generated from the same source rock.
Fig. 4. Modified Bernard diagram of natural gases from Lower Cretaceous clastic and
Upper Jurassic carbonate reservoir fields. Natural gases are generated from type III
Generally reversed carbon isotope ratios are attributed to be the
kerogen, however the location of Upper Jurassic carbonate reservoir gases are due to result of gas mixing from different sources and/or gases with
effects of TSR. different maturity (James, 1983; Jenden et al., 1993, 1988; Hao et al.,
82 H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87

Fig. 5. Comparison of compositional and isotopic data of natural data. Considering the shale gas mixture, clastic reservoir gases are primary derived gases, while due to mixture
with shale gases, carbonate reservoir gases are un-plot-able. (P: primary cracking gas, CeI: secondary cracking of oil, C-II: secondary cracking of oil and gas, C-III: secondary cracking
of gas to gas, C þ P: the mixing gas of primary cracking gas and secondary oil cracking gas).

activity is recognized by methane isotopic value ranging negative


than 50 or 60‰ (Schoell et al., 1981; Hunt, 1996; Whiticar, 1999).
Therefore the introduced gas responsible for partial reversal of
alkane carbon isotopes in both reservoirs is coal derived gas in
nature.
Mixed gases appear to have only affected the ethane in both
reservoirs, and the rest of the gas homologues do not show
noticeable change, implying that the heaviest n-alkane in the
migrated gas was ethane. With increasing in maturity, methane
and wet gas generation will increase and decrease, respectively,
implying mixing with more mature gases from the same source is
unlikely as methane isotopic values were not shifted to higher
(enriched) values (Table 2). As mentioned previously in the
geological setting section, layers of undrilled/unexplored shales
(shale gas) exist below the studied reservoirs, and unlike relative
openness in coal environment, shale gases are considered to be
closed systems. Isotopic reversals are known to be features of
closed shale systems, which low expulsion efficiency have pre-
vented the oils and gases to be expelled at their respective peak
generation. These propeties seem to be the result of in-situ mixing
and accumulation of gases generated from cracking of different
Fig. 6. “Natural gas plot” of the studied samples Natural gas plot using the Chung et al. precursors (kerogen, retained oil and wet gas) at different thermal
(1988) method. Upper Jurassic carbonate reservoir gases contain C2 gases with higher maturities (Tilley et al., 2011; Zumberge et al., 2012; Tilley and
isotopic values than Lower Cretaceous clastic, due to mixing with shale gases with
complete reversal.
Muehlenbachs, 2013).
Hao and Zou (2013) have proposed a conceptual model to
explain the procedure and the reasons for observed anomalies in
1998, 2000; Laughrey et al., 2004). Dai et al. (2004) demonstrated isotopic composition of shale gases with high thermal maturities.
that partial reversal of alkane carbon isotopes could be by following At thermal maturities lower than 1.3e1.5% Ro, gases are mainly
reasons: 1) mixing of abiogenic and biogenic gases 2) mixing of generated from kerogen cracking in a closed system, leading to the
coal-derived and oil derived gases 3) bacterial oxidation of alkane normal trend of increasing enrichment of d13C with increasing
components and 4) mixing of gases with varying thermal maturity carbon number (Lorant et al., 1998) and increases in iC4/nC4 and iC5/
from the same source or gases from different sources. The gases in nC5 ratios which exhibit maturity traits similar to gases generated
the area are proven to be thermogenic in origin and there is no in an open-system environment.
report of abiogenic gases in the area (as discussed above). Therefore Partial isotope reversal (d13C2 < d13C1 < d13C3) begins to occur as
mixing with abiogenic gases is unlikely. As mentioned above thermal maturities increases to about Ro ~1.9 (R2 represented in
ethane isotope values generated from sapropelic source rocks Fig 7). Cracking of retained oil causes an increase in ethane and
possess values lighter than 28‰ (Stahl and Carey, 1975; Dai et al., propane concentrations. The generated gases from oils are gener-
1992). The data suggest (Figs. 3 and 5, Table 2) that the d13C2 values ally enriched in d12C causing depletion in the overall d13C2 and
of introduced gases were heavier (enriched) than the gases. d13C3. However C3 gases (light isotopes) generates at higher rate,
Therefore the mixed gas should have been coal-derived gas in na- which results in higher values of (d13C2 - d13C3) with maturity (Fig
ture. The data show no sign of bacterial activity, as the bacterial 7). As thermal maturity further increases (Ro > 2.0%) apart from
H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87 83

Fig. 7. Schematic diagram depicting the transition of carbon isotope distribution patterns from normal into complete reversal through increasing thermal maturity and d13C
distribution pattern of C2 and C3 from negative divergence into positive divergence (R1: 1.3e1.5%Ro, R2: 1.9%Ro, R3 &R4: >2.0%Ro; S1: d13C1 < d13C2 < d13C3, S2: d13C2 < d13C1 < d13C3,
S3: d13C2 < d13C3 < d13C1, R4: d13C3 < d13C2 < d13C1) (Hao and Zou, 2013).

previous precursors, wet gases also start to crack and the partial however due to high concentration in both reservoirs, methane
isotopic reversal (d13C2 < d13C3 < d13C1) changes into complete were not affected by the migrated gases.
isotopic reversal (d13C3 < d13C2 < d13C1) (Fig 7, Hao and Zou, 2013 The carbon isotopic composition of the shale gases in the
and references therein). studied area below Upper Jurassic carbonate reservoir gas is highly
Considering the present depths of Upper Jurassic carbonate enriched with complete reversal, indicating progressive loss of all
reservoir, the isotopic values of the gases from these reservoirs as hydrocarbon gases due to gas loss and/or fluid rock reactions
well as their gas concentrations, the level of TSR occurring in these (Burruss and Laughrey, 2010) suggesting that deeper, economic
reservoirs and also the thermal maturities of shale gas layers, the accumulations of natural gas/shale gases deeper than Upper
authors believe that the gases in lower shale gases are completely Jurassic carbonate reservoir gas reservoir are unlikely.
reversed with thermal maturities above 2.0%Ro (Hao and Zou,
2013). 5.3. TSR occurrence in Upper Jurassic carbonate reservoir
The Closed-system of a shale gas may become more open-
system due to faulting, fracturing and/or intensive erosion of Thermochemical sulfate reduction is known to occur at tem-
overburden rocks, which can result in complete loss of free gas or peratures higher than 100e140  C (Orr, 1977; Goldstein and
even undersaturation. Loss of gases under the influence of intensive Aizenshtat, 1994; Goldhaber and Orr, 1995; Machel et al., 1995;
tectonic movements after the end of gas generation is probably the Worden et al., 1995; Machel, 2001; Mougin et al., 2007), involving
major risk of gas exploration in organic-rich shales (Hao and Zou, sulfate and organic matter/hydrocarbons. During TSR and oxidation
2013). of hydrocarbons, large amounts of H2S, CO2, CH4 and H2O are
Burial and thermal history reconstruction revealed that the area produced at different rates (Worden et al., 1995; Worden and
have experienced two noticeable uplifts (Fig. 8), the first uplift Smalley, 1996; Cai et al., 2004; Pan et al., 2006). This increase in
being the major uplift happened at Late Eocene times (37e33) the gas population leads to an increase in gas dryness coefficient
followed by the last one at 4e7 years ago (Frizon de Lamotte et al., (Zhang et al., 2008) (Table 1).
2011; Robert et al., 2014; Shabanian et al., 2009). It is likely that The concentration of the H2S gases in Lower Cretaceous clastic
during the first uplift, the deep shale gases have become less reservoir is below detection and both population and isotopic
pressurized/confined and have lost some yet adequate volume of values of CO2 gases implying no TSR occurrence in these reservoirs
the free gases which then migrated to both Upper Jurassic car- (Table 1) which is likely due to the clastic nature of the reservoir
bonate reservoir and then to Lower Cretaceous clastic reservoir. (Dai et al., 2014). On the other hand both concentrations and iso-
During the following uplift and further removal of pressure on topic values of H2S and CO2 of Upper Jurassic carbonate reservoir
shale gases, gases with higher maturity have been released, which gases, the presences of saddle and matrix dolomite, bitumens along
apparently only migrated to Upper Jurassic carbonate reservoir, with stylolite surface (Kavoosi, 2009) in the reservoir are evidences
however it is not clear if these gases have experienced TSR in Upper for TSR. As mentioned above existence of anhydrites alternative
Jurassic carbonate reservoir. with dolomite intervals in the carbonate reservoir were reported
As previously explained gases taken from Lower Cretaceous (Afshar Harb, 1979; Kavoosi, 2014; Kavoosi et al., 2009). One
clastic reservoir display no trace of secondary processes. Consid- dimensional maturity modelling in the B anticline shows that this
ering the in-reservoir mixing with shale gases, the overall con- reservoir experienced up to 160  C temperature which highlights
centrations of C2 and C3 gases are generally lower than 2% and 0.5%, the occurrence of TSR (Fig. 8).
respectively (Table 1), while overall C2 concentrations of TSR- Generally in reservoirs with sufficient amounts of sulfate, three
affected Upper Jurassic carbonate reservoir and migrated shale stage sequence described by Hao et al. (2008) occurs. Occurrence of
gases are <0.6%. This implies that the migrated shale gases con- all three stages of TSR will lead to elevated amounts of H2S and CO2
sisted of low concentrations (only C1 and C2) yet high isotopic gases, accompanied by high quantities of sulfur rich pyrobitumens
values (higher in case of gases migrated during second uplift). This in the reservoir. However in case of limited amounts of sulfate
explains why only ethane in both reservoirs show gas enrichment, concentration, TSR is restricted to condensate or at most to
84 H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87

Fig. 8. Result of one dimensional thermal maturity modelling with calibration data from the B anticline.

condensate and wet gas stage. In these reservoirs TSR mostly supported by local existence of anhydrite in Upper Jurassic car-
consumes C2þ hydrocarbon gases while leaves the methane d13C1 bonate sequences (Afshar Harb, 1979).
values unaffected (Hao et al., 2008). This causes relatively constant Theoretically, isotopically heavy CH4 and light CO2 are products
d13C1 values but varying C1/(C2þC3) values (Table 1). TSR process at of an increase in the level of TSR through hydrocarbon oxidative
Upper Jurassic carbonate reservoir is almost at the middle-end of alteration. At this stage the light CO2 is mainly derived from hy-
advanced stage (wet gas stage), and the reason for similar d13C drocarbons. On the other hand with increasing in TSR level, along
methane values in TSR affected Upper Jurassic carbonate reservoir with increasing in H2S volume the isotopic values as well as become
and Lower Cretaceous clastic reservoir with no secondary process is enriched (Aali et al., 2006). Upper Jurassic carbonate reservoir gases
due to limited concentrations of sulfates. This suggestion is possess low quantities of H2S and CO2 (Table 1), but are heavily
H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87 85

enriched in d34S and d13C values of H2S and CO2 gases (Table 2), preferentially precipitated (reacting with Mg2þ and Fe2þ). Start of
respectively; implying thermochemical sulfate reduction is not the uplift and the decrease in temperature and pressure of the reservoir
only process responsible for these values. Generally due to high have driven the reservoir fluids from saturate into sub-saturate
reactivity of H2S gases with transition metals resulting in formation state (Huang et al., 2010). The acidic fluids have dissolved the car-
of pyrite and other metal sulfides, the amount of H2S is majorly bonates and the former precipitated crystal, which in turn releases
decreases (Orr, 1977; Suleimenov and Krupp, 1994; Kotarba and CO2 gases containing heavy isotopes, mixing of these gases with
Lewan, 2013). Whereas isotopically heavy CO2 occurrence in nat- previously TSR-generated CO2 have enriched the isotopic values, as
ural gases are normally due to reaction/precipitation of lighter described by Liu et al. (2014).
member of the gase with Mg2þ, Fe2þ and Ca2þ of sulfates, through
below equation: 6. Conclusion

CO2 þ Mg2þ (Ca2þ, Fe2þ) / MgCO3 (CaCO3, FeCO3) Y þ H2O (1) Results of molecular and isotopic composition of natural gases
from Lower Cretaceous clastic reservoir and Upper Jurassic car-
Through this equation the light CO2 precipitate, while the re- bonate reservoir in the Eastern Kopeh-Dagh area indicate that:
sidual gases become enriched in d13C values.
Liu et al. (2014) have proposed that CO2 gases in natural gases 1. Gaseous hydrocarbons accumulated in both reservoirs are
with content >5% and isotopic values lower than 3‰, are pri- genetically thermogenic in nature, derived mainly from source
marily produced through TSR, while CO2 gases in natural gases with rock with kerogen type III, however enriched dD values of
content >5% but isotopic values higher (enriched) than 3‰, are methane implies kerogen type II (marine organic matter) ma-
the mixture of hydrocarbon-derived CO2 via TSR and carbonate- terial existence in the source.
dissolution CO2 by acidic fluids. 2. The Upper Jurassic carbonate reservoir natural gas samples have
As TSR proceeds through consuming heavier hydrocarbons, the higher dryness coefficient derived from TSR and gas-gas thermal
acidic gas content (H2S and CO2) increases in the reservoir. As cracking, however similar dD values of methane implies that the
mentioned previously H2S gas is highly reactive with transition mechanism responsible for the observed differences in Upper
metals and Mg2þ is considered as an active metal for reacting with Jurassic carbonate reservoir is mainly caused by TSR rather than
H2S gas (Ma et al., 2008; Machel, 2001; Worden and Smalley, 1996). thermal cracking.
Ferroan is generally presents as cement in limestones and dolo- 3. Carbon isotopic profile of both reservoirs exhibit partial reversal
stones (Liu et al., 2013; Machel, 2001). Therefore Pyrite is prefer- in ethane isotopic values, with Upper Jurassic carbonate reser-
entially precipitates during TSR in case Fe2þ exists in the voir gases possess greater reversal resulted from mixture with
environment; During TSR any FeCO3-component will preferentially highly mature completely reversed shale gases with thermal
grow into pyrite through following reaction: maturities above 2.0%Ro.
4. Molecular and isotopic values of natural gases indicate no sec-
FeCO3þ2H2O /FeS2þCO2þH2O þ H2 (2) ondary processes in Lower Cretaceous clastic reservoir, while
both concentration and isotopic values of H2S and CO2 of Upper
Since the body part of Upper Jurassic carbonate reservoir is Jurassic carbonate reservoir gases and other petrographic evi-
made of high-Mg calcite and dolomitelayers (Afshar Harb, 1979; dence in the reservoir signify TSR occurrence.
Kavoosi, 2014; Kavoosi et al., 2009), it appears that the heavily 5. d13C of carbon dioxide and heavily enriched d34S values of
enriched isotopic values of H2S gases are due to preferential reac- hydrogen sulfide in Upper Jurassic carbonate reservoir, do not
tion/precipitation of Mg2þ (since Mg2þ is considered a highly solely reflect TSR. The enriched isotopic value of CO2 is resulted
reactive transition metal with H2S; Worden and Smalley, 1996; from mixing of previously TSR produced CO2 and CO2 produced
Machel, 2001; Ma et al., 2008), and Fe2þ with H2S gases. These through carbonate rock dissolution by acidic gases. While
reaction/precipitation of lighter isotopes (d12C, d32S) will make the enriched d34S values are due to reaction/precipitation of light
residual gases to become enriched in heavier isotopes (d13C, d34S). H2S gases with Mg2þ and Fe2þ in the reservoir.
As mentioned before by burial and thermal model reconstruc- 6. Burial reconstruction model shows that the area has experi-
tion of both gas reservoirs (Fig. 8), the area have experience two enced two noticeable uplifts. Prior to these uplifts, Upper
uplifts. Prior to these uplifts Upper Jurassic carbonate reservoir Jurassic carbonate reservoir have experienced temperatures
have experienced temperatures far greater (up to 160  C temper- optimum for thermochemical sulfate reduction, however the
ature) than the proposed threshold for TSR process (100e140  C). following uplifts have slowed if not stopped the rate of TSR.
During this period TSR at Upper Jurassic carbonate reservoir have 7. It is likely that during the first major uplift occurrence (33e37
proceeded and reached the end of second stage (wet gas stage), as million years ago), released gases from shale gases have
described by Hao et al. (2008). During the wet gas level, depleted C2 migrated to both reservoirs resulting in the observed reversal
and C3 gases generated during thermal cracking of retained oils enrichment, however during subsequent uplift released gases
(Hao and Zou, 2013; Prinzhofer and Huc, 1995) were consumed, from shale gases only migrated to Upper Jurassic carbonate
which could be the reason for similar values of C3 gases in both reservoir.
reservoirs (consumption of light-methane gases generated previ- 8. Progressive loss of hydrocarbon gases due to gas loss and fluid
ously during thermal cracking and previous levels of TSR at the end rock reactions in shale gases have led to high enrichment and
of wet gas level, could be the reason for lower methane concen- complete reversal of carbon isotopic composition of the shale
tration in Upper Jurassic carbonate reservoir). gases in the studied area, suggesting that economic accumula-
However it appears that the start of uplift have decreased the tions of natural gas/shale gases deeper than Upper Jurassic
temperature, therefore slowed if not stopped the process of n- carbonate reservoir are unlikely.
alkane cracking and H2S production, as higher temperatures are
needed to cleave the bonds between carbons with decreasing in Acknowledgments
carbon number (Curt Wentrup, 1984). As mentioned before TSR
elevates the H2S and CO2 content in the reservoir, after saturating The authors would like to extend their thanks to the Exploration
the reservoir fluids, the lighter isotopes of carbon and sulfur are Directorate of the National Iranian Oil Company (NIOC) for
86 H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87

providing the samples and the financial support of the project. The Sichuan Basin, China. Am. Assoc. Pet. Geol. Bull. 92, 611e637.
Hao, F., Li, S., Gong, Z., Yang, J., 2000. Thermal regime, interreservoir compositional
invaluable help of the anonymous reviewers of this manuscript,
heterogeneities, and reservoir-filling history of the Dongfang Gas Field, Ying-
whose ideas and comments improved the manuscript, are also gehai Basin, South China Sea: evidence for episodic fluid injections in over-
acknowledged. We wish to thank M. Soltani and as well as Asso- pressured basins? Am. Assoc. Pet. Geol. Bull. 84, 607e626.
ciated Editor of JMPG Dr Daniel S Alessi, for his invaluable help and Hao, F., Li, S., Sun, Y., Zhang, Q., 1998. Geology, compositional heterogeneities, and
geochemical origin of the Yacheng gas field, Qiongdongnan Basin, South China
reviewing thoroughly the whole manuscript. sea. Am. Assoc. Pet. Geol. Bull. 82, 1372e1384.
Hao, F., Zou, H., 2013. Cause of shale gas geochemical anomalies and mechanisms
for gas enrichment and depletion in high-maturity shales. Mar. Pet. Geol. 44,
References 1e12.
Huang, S., Huang, K., Tong, H., Liu, L., Sun, W., Zhong, Q., 2010. Origin of CO2 in
Aali, J., Rahimpour-Bonab, H., Kamali, M.R., 2006. Geochemistry and origin of the natural gas from the triassic feixianguan formation of northeast sichuan basin.
world's largest gas field from Persian Gulf, Iran. J. Pet. Sci. Eng. 50, 161e175. Sci. China Earth Sci. 53, 642e648.
Adabi, M.H., 2009. Multistage dolomitization of upper jurassic mozduran formation, Hunt, J.M., 1996. Petroleum Geology and Geochemistry (Free. New York).
Kopet-Dagh Basin, ne Iran. Carbonates Evaporites 24, 16e32. Isaksen, G.H., Khalylov, M., 2007. Controls on Hydrogen Sulfide Formation in a
Afshar Harb, A., 1979. The Stratigraphy, Tectonics and Petroleum Geology of the Jurassic Carbonate Play, Turkmenistan.
Kopet-dogh Region, Northern Iran (Unpubl). James, A.T., 1983. Correlation of natural gas by use of carbon isotopic distribution
Bernard, B.B., Brooks, J.M., Sackett, W.M., 1978. Light hydrocarbons in recent Texas between hydrocarbon components. Am. Assoc. Pet. Geol. Bull. 67, 1176e1191.
continental shelf and slope sediments. J. Geophys. Res. Oceans 83, 4053e4061. Jenden, P.D., Kaplan, I.R., Hilton, D.R., Craig, H., 1993. Abiogenic hydrocarbons and
Bernard, B.B., Brooks, J.M., Sackett, W.M., 1976. Natural gas seepage in the Gulf of mantle helium in oil and gas fields. U. S. Geol. Surv. Prof. Pap. States 1570.
Mexico. Earth Planet. Sci. Lett. 31, 48e54. Jenden, P.D., Kaplan, I.R., Poreda, R.J., Craig, H., 1988. Origin of nitrogen-rich gases in
Brookfield, M.E., Hashmat, A., 2001. The geology and petroleum potential of the the Californian Great Valley : evidence from helium, carbon and nitrogen
North Afghan platform and adjacent areas (northern Afghanistan, with isotope ratios. Geochim. Cosmochim. Acta 52, 851e861.
parts of southern Turkmenistan, Uzbekistan and Tajikistan). Earth Sci. Rev. 55, Kavoosi, M.A., 2014. Inorganic control on original carbonate mineralogy and crea-
41e71. tion of gas reservoir of the Upper Jurassic carbonates in the Kopet-Dagh Basin,
Burruss, R.C., Laughrey, C.D., 2010. Carbon and hydrogen isotopic reversals in deep NE, Iran. Carbonates Evaporites 29, 419e432.
basin gas: evidence for limits to the stability of hydrocarbons. Org. Geochem. 41, Kavoosi, M.A., 2009. Sedimentary Environments and Sequence Stratigraphy of the
1285e1296. Upper Jurassic Mozduran Formation in the Eastern and Central Areas of the
Cai, C., Xie, Z., Worden, R.H., Hu, G., Wang, L., He, H., 2004. Methane-dominated Kopet-dagh Basin, NE Iran (Unpublished Ph.D Thesis). Teacher Training Uni-
thermochemical sulphate reduction in the triassic feixianguan formation east versity of Tehran, pp. 1e175 (in Persian, with English abstract).
sichuan basin, China: towards prediction of fatal H2S concentrations. Mar. Pet. Kavoosi, M.A., Lasemi, Y., Sherkati, S., Moussavi-Harami, R., 2009. Facies analysis and
Geol. 21, 1265e1279. depositional sequences of the upper jurassic mozduran formation, a carbonate
Chen, A.D., 1994. Origin and migration of natural gas in Ordovician reservoir in Shan reservoir in the kopet dagh basin, NE Iran. J. Pet. Geol. 32, 235e260.
Gan Ning Basin central gas field. Acta Pet. Sin. 15, 1e10. Kotarba, M.J., Lewan, M.D., 2013. Sources of natural gases in Middle Cambrian
Chung, H.M., Gormly, J.R., Squires, R.M., 1988. Origin of gaseous hydrocarbons in reservoirs in Polish and Lithuanian Baltic Basin as determined by stable isotopes
subsurface environments: theoretical considerations of carbon isotope distri- and hydrous pyrolysis of Lower Palaeozoic source rocks. Chem. Geol. 345,
bution. Chem. Geol. 71, 97e104. 62e76.
Cody, J.D., Hutcheon, I.E., 1994. Regional water and gas geochemistry of the Kuleshov, A.V., Ignatova, V.A., 1990. Present temperature regime of Jurassic and
Mannville Group and associated horizons, southern Alberta. Bull. Can. Pet. Geol. Lower Cretaceous rocks of the Amu-Darya basin as related to their oil and gas
42, 449e464. productivity. In: Kleschev, K.A., Shein, V.S. (Eds.), Geology and Geodynamics of
Dai, J., Gong, D., Ni, Y., Huang, S., Wu, W., 2014. Stable carbon isotopes of coal- Petroleum Basins of the USSR [Geologiya I Geodinamika Neftegazonosnykh
derived gases sourced from the Mesozoic coal measures in China. Org. Geo- Basseynov SSSR]. VNIGNI, Moscow, pp. 158e165.
chem. 74, 123e142. http://dx.doi.org/10.1016/j.orggeochem.2014.04.002. Laughrey, C.D., Billman, D.A., Canich, M.R., 2004. Petroleum geology and
Curt Wentrup, 1984. Reactive Molecules, the Natural Reactive Intermediates in geochemistry of the Council Run gas field, north central Pennsylvania. Am.
Organic Chemistry, Curt Wentrup, Department of Chemistry, University of Assoc. Pet. Geol. Bull. 88, 213e239.
Marburg West Germany. John Wiley & Sons. Liu, Q., Jin, Z., Meng, Q., Wu, X., Jia, H., 2015. Genetic types of natural gas and filling
Dai, J., Li, J., Luo, X., Zhang, W., Hu, G., Ma, C., Guo, J., Ge, S., 2005. Stable carbon patterns in Daniudi gas field, Ordos Basin, China. J. Asian Earth Sci. 107, 1e11.
isotope compositions and source rock geochemistry of the giant gas accumu- http://dx.doi.org/10.1016/j.jseaes.2015.04.001.
lations in the Ordos Basin, China. Org. Geochem. 36, 1617e1635. Liu, Q., Jin, Z., Wu, X., Liu, W., Gao, B., Zhang, D., Li, J., Hu, A., 2014. Origin and carbon
Dai, J., Xia, X., Qin, S., Zhao, J., 2004. Origins of partially reversed alkane d13 C values isotope fractionation of CO2 in marine sour gas reservoirs in the Eastern
for biogenic gases in China. Org. Geochem. 35, 405e411. Sichuan Basin. Org. Geochem 74, 22e32. http://dx.doi.org/10.1016/
Dai, J.X., Pei, X.G., Qi, H.F., 1992. Natural Gas Geology in China, vol. 1. j.orggeochem.2014.01.012.
Deines, P., 1980. The isotopic composition of reduced organic carbon. Handb. En- Liu, Q.Y., Dai, J.X., Zhang, T.W., Li, J., Qin, S.F., Liu, W.H., 2008. Genetic types of natural
viron. Isot. Geochem. 329e406. gas and their distribution in Tarim Basin, NW China. Nat. Sci. Sustain. Technol.
Draper, J.J., Boreham, C.J., 2006. Geological controls on exploitable coal seam gas Res. Prog. 1, 247e268.
distribution in queensland. Appea J. 46, 343e366. http://dx.doi.org/10.2473/ Liu, Q.Y., Worden, R.H., Jin, Z.J., Liu, W.H., Li, J., Gao, B., Zhang, D.W., Hu, A.P., Yang, C.,
shigentosozai.117.923. 2013. TSR versus non-TSR processes and their impact on gas geochemistry and
Faiz, M., Hendry, P., 2006. Significance of microbial activity in Australian coal bed carbon stable isotopes in Carboniferous, Permian and Lower Triassic marine
methane reservoirsda review. Bull. Can. Pet. Geol. 54, 261e272. carbonate gas reservoirs in the Eastern Sichuan Basin, China. Geochim. Cos-
Flores, R.M., Rice, C.A., Stricker, G.D., Warden, A., Ellis, M.S., 2008. Methanogenic mochim. Acta 100, 96e115.
pathways of coal-bed gas in the Powder River Basin, United States: the geologic Lorant, F., Prinzhofer, A., Behar, F., Huc, A.-Y., 1998. Carbon isotopic and molecular
factor. Int. J. Coal Geol. 76, 52e75. constraints on the formation and the expulsion of thermogenic hydrocarbon
Frizon de Lamotte, D., Raulin, C., Mouchot, N., Wrobel-Daveau, J., Blanpied, C., gases. Chem. Geol. 147, 249e264.
Ringenbach, J., 2011. The southernmost margin of the Tethys realm during the Ma, Y., Zhang, S., Guo, T., Zhu, G., Cai, X., Li, M., 2008. Petroleum geology of the
Mesozoic and Cenozoic: initial geometry and timing of the inversion processes. Puguang sour gas field in the Sichuan Basin, SW China. Mar. Pet. Geol. 25,
Tectonics 30. 357e370.
Fürsich, F.T., Wilmsen, M., Seyed-Emami, K., Majidifard, M.R., 2009. The mid- Machel, H.G., 2001. Bacterial and thermochemical sulfate reduction in diagenetic
cimmerian tectonic event (bajocian) in the alborz mountains, northern Iran: settingsdold and new insights. Sediment. Geol. 140, 143e175.
evidence of the break-up unconformity of the South Caspian basin. Geol. Soc. Machel, H.G., Krouse, H.R., Sassen, R., 1995. Products and distinguishing criteria of
Lond. Spec. Publ. 312, 189e203. bacterial and thermochemical sulfate reduction. Appl. Geochem. 10,
Galimov, E.M., 1988. Sources and mechanisms of formation of gaseous hydrocar- 373e389.
bons in sedimentary rocks. Chem. Geol. 71, 77e95. Mortazavi, M., Moussavi-Harami, R., Mahboubi, A., Nadjafi, M., 2014. Geochemistry
Goldhaber, M.B., Orr, W.L., 1995. Kinetic controls on thermochemical sulfate of the late jurassiceearly cretaceous shales (shurijeh formation) in the intra-
reduction as a source of sedimentary H2S. In: ACS Symposium Series. American continental Kopet-Dagh basin, northeastern Iran: implication for provenance,
Chemical Society, Washington, DC, pp. 412e425 (1974). source weathering, and paleoenvironments. Arab. J. Geosci. 7, 5353e5366.
Goldstein, T.P., Aizenshtat, Z., 1994. Thermochemical sulfate reduction a review. Mougin, P., Lamoureux-Var, V., Bariteau, A., Huc, A.Y., 2007. Thermodynamic of
J. Therm. Anal. 42, 241e290. thermochemical sulphate reduction. J. Pet. Sci. Eng. 58, 413e427.
Golonka, J., 2004. Plate tectonic evolution of the southern margin of Eurasia in the Moussavi-Harami, R., Brenner, R.L., 1993. Diagenesis of non-marine petroleum
Mesozoic and Cenozoic. Tectonophysics 381, 235e273. reservoirs: the neocomian (lower cretaceous) Shurijeh Formation, Kopet-Dagh
Guoyi, H., Jin, L., Xiuqin, S., Zhongxi, H., 2010. The origin of natural gas and the basin, NE Iran. J. Pet. Geol. 16, 55e72.
hydrocarbon charging history of the Yulin gas field in the Ordos Basin, China. Moussavi-Harami, R., Brenner, R.L., 1992. Geohistory analysis and petroleum
Int. J. Coal Geol. 81, 381e391. reservoir characteristics of lower cretaceous (neocomian) sandstones, eastern
Hao, F., Guo, T., Zhu, Y., Cai, X., Zou, H., Li, P., 2008. Evidence for multiple stages of oil Kopet-Dagh basin, northeastern Iran (1). Am. Assoc. Pet. Geol. Bull. 76,
cracking and thermochemical sulfate reduction in the Puguang gas field, 1200e1208.
H. Saadati et al. / Marine and Petroleum Geology 78 (2016) 76e87 87

Orr, W.L., 1977. Geologic and geochemical controls on the distribution of hydrogen Turich, C., Ashby, M., 2011. Biogeochemistry of Microbial Coal-bed Methane.
sulfide in natural gas. Adv. Org. Geochem. 1975, 571e597. Stra˛ poc, D., Mastalerz, M., Schimmelmann, A., Drobniak, A., Hedges, S., 2008.
Osborn, S.G., McIntosh, J.C., 2010. Chemical and isotopic tracers of the contribution Variability of geochemical properties in a microbially dominated coalbed gas
of microbial gas in Devonian organic-rich shales and reservoir sandstones, system from the eastern margin of the Illinois Basin, USA. Int. J. Coal Geol. 76,
northern Appalachian Basin. Appl. Geochem. 25, 456e471. 98e110.
Pan, C., Yu, L., Liu, J., Fu, J., 2006. Chemical and carbon isotopic fractionations of Suleimenov, O.M., Krupp, R.E., 1994. Solubility of hydrogen sulfide in pure water and
gaseous hydrocarbons during abiogenic oxidation. Earth Planet. Sci. Lett. 246, in NaCl solutions, from 20 to 320 C and at saturation pressures. Geochim.
70e89. Cosmochim. Acta 58, 2433e2444.
Poursoltani, M.R., Gibling, M.R., 2011. Composition, porosity, and reservoir potential Tilley, B., McLellan, S., Hiebert, S., Quartero, B., Veilleux, B., Muehlenbachs, K., 2011.
of the middle jurassic Kashafrud Formation, northeast Iran. Mar. Pet. Geol. 28, Gas isotope reversals in fractured gas reservoirs of the western Canadian
1094e1110. Foothills: mature shale gases in disguise. Am. Assoc. Pet. Geol. Bull. 95,
Prinzhofer, A.A., Huc, A.Y., 1995. Genetic and post-genetic molecular and isotopic 1399e1422.
fractionations in natural gases. Chem. Geol. 126, 281e290. Tilley, B., Muehlenbachs, K., 2013. Isotope reversals and universal stages and trends
Robert, A.M.M., Letouzey, J., Kavoosi, M.A., Sherkati, S., Müller, C., Verge s, J., of gas maturation in sealed, self-contained petroleum systems. Chem. Geol. 339,
Aghababaei, A., 2014. Structural evolution of the Kopeh Dagh fold-and-thrust 194e204.
belt (NE Iran) and interactions with the South Caspian Sea basin and Amu Tissot, B., Welte, D., 2012. Petroleum Formation and Occurrence: a New Approach to
Darya basin. Mar. Pet. Geol. 57, 68e87. Oil and Gas Exploration. Springer Science & Business Media.
Schoell, M., 1980. The hydrogen and carbon isotopic composition of methane from Ulmishek, G.F., 2004. Petroleum geology and resources of the Amu-Darya basin,
natural gases of various origins. Geochim. Cosmochim. Acta 44, 649e661. Turkmenistan, Uzbekistan, Afghanistan, and Iran. USGS Bull. 38.
http://dx.doi.org/10.1016/0016-7037(80)90155-6. USGS, U.S., 2000. Geological Survey World Petroleum Assessment 2000-description
Schoell, M., Teschner, M., Wehner, H., Durand, B., Oudin, J.L., 1981. Maturity related and Results. USGS Digit. Data Ser. DDS-60.
biomarker and stable isotope variations and their application to oil/source rock Whiticar, M.J., 1999. Carbon and hydrogen isotope systematics of bacterial forma-
correlation in the Mahakam Delta, Kalimantan. Adv. Org. Geochem. 10, tion and oxidation of methane. Chem. Geol. 161, 291e314.
156e163. Whiticar, M.J., 1996. Stable isotope geochemistry of coals, humic kerogens and
Semenovich, V.V., Maksimov, S.P., Pankina, R.G., Mekhtieva, V.L., Gurieva, S.M., 1983. related natural gases. Int. J. Coal Geol. 32, 191e215.
Genesis of Hydrogen Sulfide of the Dauletabad-donmez Field: Geologiya Nefti I Worden, R.H., Smalley, P.C., 1996. H2S-producing reactions in deep carbonate gas
Gaza, No. 6. reservoirs: khuff Formation, Abu Dhabi. Chem. Geol. 133, 157e171.
Shabanian, E., Bellier, O., Siame, L., Arnaud, N., Abbassi, M.R., Cocheme , J., 2009. New Worden, R.H., Smalley, P.C., Oxtoby, N.H., 1995. Gas souring by thermochemical
tectonic configuration in NE Iran: active strike-slip faulting between the Kopeh sulfate reduction at 140 C. Am. Assoc. Pet. Geol. Bull. 79, 854e863.
Dagh and Binalud mountains. Tectonics 28. Yongchang, X., Ping, S., 1996. A study of natural gas origins in China. Am. Assoc. Pet.
Silverman, S.R., 1971. Influence of petroleum origin and transformation on its dis- Geol. Bull. 80, 1604e1614.
tribution and redistribution in sedimentary rocks. In: 8th World Petroleum Zanchetta, S., Berra, F., Zanchi, A., Bergomi, M., Caridroit, M., Nicora, A.,
Congress. World Petroleum Congress. Heidarzadeh, G., 2013. The record of the Late Palaeozoic active margin of the
Smith, J.W., Pallasser, R.J., 1996. Microbial origin of Australian coalbed methane. Am. Palaeotethys in NE Iran: constraints on the Cimmerian orogeny. Gondwana Res.
Assoc. Pet. Geol. Bull. 80, 891e897. 24, 1237e1266.
Sokolova, I.M., Abryutina, N.I., Makarov, V.V., Kuldzhaev, B.A., Rusinova, G.V., Zhang, T., Amrani, A., Ellis, G.S., Ma, Q., Tang, Y., 2008. Experimental investigation on
Petrov, A.A., 1993. Biomarkers in Gas Condensates of Eastern Turkmenistan: thermochemical sulfate reduction by H 2 S initiation. Geochim. Cosmochim.
Geokhimiya, No. 1. Acta 72, 3518e3530.
Stahl, W.J., Carey, B.D., 1975. Source-rock identification by isotope analyses of nat- Zumberge, J., Ferworn, K., Brown, S., 2012. Isotopic reversal (“rollover”) in shale
ural gases from fields in the Val Verde and Delaware basins, west Texas. Chem. gases produced from the Mississippian Barnett and Fayetteville formations.
Geol. 16, 257e267. Mar. Pet. Geol. 31, 43e52.
Stra˛ poc, D., Mastalerz, M., Dawson, K., Macalady, J., Callaghan, A.V., Wawrik, B.,

You might also like