You are on page 1of 11

International Journal of Coal Geology 115 (2013) 4151

Contents lists available at SciVerse ScienceDirect

International Journal of Coal Geology


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

Organic geochemistry and elements distribution in Dahuangshan oil


shale, southern Junggar Basin: Origin of organic matter and
depositional environment
Shu Tao a,b,, Dazhen Tang a, Hao Xu a, Jianlong Liang a, Xuefeng Shi c
a
b
c

Coal Reservoir Laboratory of National CBM Engineering Center, School of Energy Resources, China University of Geosciences (Beijing), Beijing 100083, China
Coalbed Methane Resources and Reservoir Formation Process Key Laboratory of Ministry of Education, China University of Mining and Technology, Xuzhou 221116, China
CNOOC Energy Technology & Services-Oileld Engieering Research Institute, Tianjin 300452, China

a r t i c l e

i n f o

Article history:
Received 3 March 2013
Received in revised form 14 May 2013
Accepted 16 May 2013
Available online 25 May 2013
Keywords:
Geochemistry
Rare earth elements
Paleoenvironment
Oil shale
Dahuangshan

a b s t r a c t
The Dahuangshan oil shale, located in the northern Bogda Mountain, on the southern margin of the Junggar
Basin, was deposited in a Late Permian lacustrine environment. A combined investigation of element and
organic geochemistry was performed to dene the source rock potential, the paleoenvironment, and source
of the organic matter. Thick sequences of oil shales with an average thickness of 638 m were deposited in
Lucaogou Formation which mainly consists of oil shale, argillaceous dolomite, silty claystone, tuff, limestone,
and dolomitic marl. A spot of plant stem fossils and abundance of pyrite crystals, shtail and sh skeleton can
also be found there.
Analyzed oil shale samples from Dahuangshan area are characterized by high total organic carbon (TOC) contents (5.634.75%), S2 (22.65199.25 mg HC/g rock), hydrogen index (HI, 3591068 mg HC/g TOC), and oil
yield (4.926.6%), indicating the oil shales have excellent source rock potential. Tmax values (433453 C)
show an early to medium maturation stage of organic matter, which is supported by organic geochemical
maturation parameters. All of the obtained kerogen types are types II and I, with oil prone source rock potential.
Dahuangshan oil shale samples are rich in SiO2 (68.59%), followed by Al2O3 (10.18%) and Fe2O3 (5.43%). Compared with average shale and North American Shale Composite (NASC), analyzed oil shale samples are obviously
enriched in P (0.71%). There is a signicant correlation between Al2O3 and Fe2O3, MgO, K2O, MnO, Cu, Ba, Co, and
Ni for their association with clay minerals. Besides, the signicant correlations between Fe2O3 and MnO, Co, and
Ni are considered to result from their similarity on geochemical behavior. All selected oil shales are characterized
by distinctly sloping light rare earth elements (LREE) trends (LaN/SmN = 2.705.95) accompanied by at heavy
rare earth elements (HREE) trends, with distinct Eu negative anomalies (0.600.73). Two slightly different patterns of REEs in the oil shale samples are distinguished by the difference in Ce depletion and Nd anomaly.
In addition, Dahuangshan oil shale samples are characterized by short- to middle-chain n-alkanes, low carbon
preference index (CPI) values (0.931.24), single peak composed of nC20 or nC22, low Pr/Ph (0.410.91), relatively
high Homohop index (0.0610.99), and high concentrations of C27 sterane, indicating reducing, deep-water, and
moderate saline environment with prevalent contribution of algae and microorganisms to organic matter
accumulation.
2013 Elsevier B.V. All rights reserved.

1. Introduction
With rapid increases in consumption of energy and chemicals, oil
supply and demand imbalances are intensifying so as to become
a restraining factor on economic growth in China. China had imported
about 179 million (Fu et al., 2010) tons of crude oil in 2008 and over
250 million tons in 2011. Oil shale, one of the substantial unconventional
Corresponding author at: Coal Reservoir Laboratory of National CBM Engineering
Center; School of Energy Resources, China University of Geosciences (Beijing), Beijing
100083, China. Tel.: + 86 10 82322011.
E-mail address: peach888@163.com (S. Tao).
0166-5162/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.coal.2013.05.004

fossil resources that can be produced and converted to liquid fuels, has
received increasing attention. The Third Chinese Oil and Gas Resource Assessment and the 2007 World Energy Survey showed that a total oil shale
resource of some 720 Gt is located across 22 provinces, 47 basins, and 80
deposits. The shale oil resource has been estimated at some 48 Gt, which
is highly signicant for alleviating the pressure of petroleum supplies (Liu
et al., 2007; WEC, 2007). At present, retorting and combustion for power
generation are the main patterns of oil shale application. In 2011, shale
oil production by retorting technology was about 1.46 million tons all
over the world, of which about 650,000 t were produced in China, including 7 oil shale retorting plants located in 5 provinces (Li, 2012). Estonia,
the biggest electricity producer from oil shale in the world (Hamburg,

42

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

2011), whose total generating capacities reached up to 3200 MW with


about 15 million tons of oil shale was used for power generation in 2011.
Thick sequences of organic-rich lacustrine oil shales have been
reported to underlie much of the Junggar Basin in Xinjiang Uygur
Autonomous Region, northwest China. Several authors have ranked
these organic-rich lacustrine mudstones (oil shales) among the thickest
and richest petroleum source rock intervals in the world (e.g. Carroll
et al., 1992; Graham et al., 1990; Watson et al., 1987). Previous researches
have focused on the oil yield, deposition, development, resources, and
metallogenic characteristics of oil shale in this area (e.g. Tao et al., 2010,
2011, 2012a,b,c). Until now, however, no available publications have
addressed the geochemical characteristics of oil shale in this area.
In the current study, the petroleum potential and the thermal maturity of the organic matter contained in Dahuangshan oil shales from
the southern Junggar Basin were studied by Rock-Eval pyrolysis and
some biomarker parameters; the occurrence and distribution of the
major and trace elements in the oil shales were studied in order to determine the geochemical background of this basin; the sedimentary and
organic geochemical characteristics of selected oil shale samples were
examined to discuss the source of organic matter and paleoenvironment
of the Dahuangshan oil shales.

2. Geological setting
Bogda Mountain is situated in the eastern part of Tianshan Mountain
range and is located on the southern margin of the Junggar Basin which
is a large, organic-rich foreland basin in northwest China (Jiao et al.,
2007; Tao et al., 2012a). Thick sequences of organic-rich lacustrine oil
shales are exposed in the foothills of Bogda Mountain (Carroll, 1998),
including 13 different oil shale mining areas, and eight of them have
been studied by us in recent years (Fig. 1A) (Tao et al., 2010, 2011,
2012a,b,c). Carroll et al. (1992) documented three Upper formations
that contain organic-rich mudstones. From oldest to youngest, they are
the Jingjingzigou, Lucaogou, and Hongyanchi formations, among which
extremely rich and oil-prone oil shales are discovered in Lucaogou
Formation.
The Dahuangshan region is located in the eastern part of Bogda
Mountain oil shale belt (Fig. 1A). The Permian Jingjingzigou, Lucaogou,
and Wutonggou formations are the major outcropping seams in this
area (Fig. 1B). The Lucaogou Formation consists of a sequence of dolomitic mudstone, dolomitic marl, argillaceous dolomite, limestone, silty
claystone, tuff, and oil shale (Fig. 2). In Dahuangshan area, the thickness
of the Lucaogou Formation (average 845 m) is larger than that of other

Fig. 1. (A) Simplied map showing geological setting of northern Bogda Mountain, and the location of the study area. (B) Simplied geological map of the Dahuangshan oil shales,
showing three oil shale proles, and two boreholes.

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

43

3. Samples and analytical procedures


Weathering is known to affect amount and quality of organic matter
in petroleum source rocks (Clayton and King, 1987; Leythaeuser, 1973).
Littke et al. (1991) noted that pyrite, sulfur, and organic carbon content
were altered by weathering. Therefore, the prole samples were collected after digging about 40 cm into the rock to minimize the effects
of surface weathering. All samples were carefully packed and then immediately sent to the laboratory for experiments.
A total of 42 outcrop oil shale samples and some interbedded
rocks were collected from section No.2 (Fig. 2). All of the oil shale
samples were selected for total organic carbon (TOC), Rock-Eval pyrolysis,
ash yield, total sulfur, organic sulfur, and Gray-King low-temperature dry
distillation analyses. Then ten of them were analyzed by X-ray uorescence (XRF), inductively-coupled plasma mass spectrometer (ICP-MS),
gas chromatography (GC) and gas chromatographymass spectrometry
(GCMS).
The samples for geochemical analysis were all crushed and ground
to less than 200 mesh. TOC and organic sulfur values were determined in the Geological Laboratory of Exploration and Development
Research Institute of PetroChina Huabei Oileld Company, following
the Chinese National Standard methods GB/T 19145-2003 and GB/T
215-2003, respectively. Rock-Eval pyrolysis data were performed on
a Rock-Eval II instrument following the guidelines established by
Espitali et al. (1985). The samples were analyzed in the Petroleum
Geology Research Center, China Petroleum Exploration and Development Research Institute. Ash yield, total sulfur, and Gray-King lowtemperature dry distillation were conducted at the Xinjiang Institute
of Coal Science and Coal Testing Laboratory, following the Chinese
National Standard methods GB/T212-2001, GB/T214-2007, and GB/T
1341-2001, respectively. The XRF was used to determine the oxides
of major elements, including SiO2, Al2O3, CaO, K2O, Na2O, Fe2O3,
MnO, MgO, TiO2, and P2O5. Trace elements were determined by an Element 6000 inductively-coupled plasma mass spectrometer (ICP-MS).
Both of the tests were determined in the Beijing Research Institute
of Uranium Geology, following the method described by Ryu et al.
(2011). The extraction and extracts separation were performed by the
method of Bolou-Bi et al. (2010). GC of the saturated hydrocarbon fraction was examined on an Agilent 7890 with a quartz capillary column
(30 m 0.25 mm 0.25 m lm thickness). GCMS of saturated hydrocarbon fractions was acquired on an Agilent7890-5975c instrument
(tted with a HP-5MS quartz capillary column of 60-m length, 0.25-mm
inner diameter, and 0.25-m lm thickness). Helium was used as the
carrier gas. The oven was held for 1 min at 50 C, programmed from
50 C to 120 C at 20 C/min, and then 120 C to 250 C at 4 C/min,
nally 250 C to 310 C at 3 C/min, with a nal holding time of
30 min at 310 C. The selected ion monitoring capabilities of the data
acquisition system permitted specic ions to be monitored, such as
n-alkanes (m/z 85), tricyclic terpanes and hopanes (m/z 191), and
steranes (m/z 217) (Amijaya et al., 2006; Korkmaz and Glbay, 2007).
Fig. 2. Stratigraphic column for the Lucaogou Formation in Dahuangshan region.

4. Results and discussion


4.1. Rock-Eval pyrolysis and TOC
areas (568 m; Tao et al., 2012a), and the thickness of total oil shale sequence is up to 638 m.
Samples for the present study were obtained from three measured
outcrop sections and two boreholes (Fig. 1B). The oil shale layers of
each prole are well exposed. Two types of facies sequences are developed in the Dahuangshan sections. Dark-black oil shale interbedded by
argillaceous dolomite, silty claystone, tuff, limestone, dolomitic marl is
developed in the middleupper part of the section (b650 m), containing
a spot of plant stem fossils, and an abundance of shtail and sh skeleton.
The thickness of oil shale is obviously thinning out in the lower part
(>650 m), and it occurs as thin layers sandwiched among surrounding
rocks (Fig. 2).

Rock-Eval pyrolysis is used to determine the petroleum potential,


thermal maturity of the organic matter and its ability to generate oil
and/or gas (Alaug, 2011). The pyrolysis gives rise parameters as S1,
S2, S3, hydrogen index (HI), oxygen index (OI), production index
(PI) and Tmax.
The Rock-Eval and TOC data are summarized in Table 1. The TOC
content of 42 oil shale samples ranges from 5.6% to 34.75 wt.%
(mean 13.79 wt.%). Rock-Eval S1 and S2 are 0.293.29 and 22.65
199.25 mg HC/g rock, respectively. The HI differs widely from 359 to
1068 mg HC/g TOC. High TOC contents, with high S2 and HI values indicate that Dahuangshan oil shales have excellent source rock potential

44

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

(Littke et al., 1998; Peters and Cassa, 1994; Tissot and Welte, 1984). The
extremely low PI-values (0.0040.037) indicate immature organic
matter whereas elevated Tmax values (433453 C) (Table 1) imply an
early to medium maturation stage.

4.2. Kerogen types


The graph of S2 vs. TOC is used to indicate the kerogen type
present and its hydrocarbon potential (e.g. Clayton and Ryder, 1984;
Cooper and Barnard, 1984; Cornford et al., 1998; Dahl et al., 2004;
Demaison and Moore, 1980; Langford and Blanc-Valleron, 1990). As
shown in Fig. 3, all of the obtained kerogen types are type II and I,
with a good potential of oil generation.
The kerogen quality and maturity are determined by plotting HI
versus Tmax rather than HI versus OI (Fig. 4; Alaug, 2011). The kerogen
type designations are entirely based on the HI (Hunt, 1996). As shown
in Fig. 4, the studied samples are at early to medium mature stage

with moderate to high HI, which ts well with the predominance of


oil prone source rock potential.
4.3. Major element geochemistry
Major element data in conjunction with mineralogical data may
be used to establish the elementmineral associations for oil shales.
Although the element associations may vary from one oil shale to another, a correlation analysis would demonstrate the general trends
(Fu et al., 2010).
As shown in Table 2, the organic sulfur (So,d) content of Dahuangshan
oil shale samples varies between 0.01 and 0.26%. The samples have a
high ash yield (5176%), with low total (St,d) contents (0.310.89%)
and relatively high oil yield (4.926.6%). SiO2 is the dominant constituent with an average of 68.59%, next come Al2O3 (6.3712.88%) and
Fe2O3 (4.326.84%).
The content of SiO2, CaO, Fe2O3, and Na2O is close to the average value
of average shale (Clarke, 1924) and North American Shale Composite

Table 1
Results of Rock-Eval/TOC analysis and calculated parameters.
Sample no.

TOCa
(wt.%)

S1b
(mg HC/g)

S2 c
(mg HC/g)

S3 d
(mg CO2/g)

HIe
(mg HCi/g TOC)

OIf
(mg CO2/g TOC)

PIg
(S1 / S1 + S2)

Tmaxh
(C)

O-01
O-02
O-03
O-04
O-05
O-06
O-07
O-08
O-09
O-10
O-11
O-12
O-13
O-14
O-15
O-16
O-17
O-18
O-19
O-20
O-21
O-22
O-23
O-24
O-25
O-26
O-27
O-28
O-29
O-30
O-31
O-32
O-33
O-34
O-35
O-36
O-37
O-38
O-39
O-40
O-41
O-42

9.71
7.45
13.20
21.12
13.90
5.86
11.00
9.12
7.86
13.92
31.17
12.56
17.77
12.35
9.35
8.51
7.67
15.11
10.16
17.81
29.02
6.01
14.39
9.38
6.69
11.41
5.60
27.06
31.99
9.30
11.85
32.23
7.13
10.61
15.69
7.44
10.22
13.33
34.75
6.31
9.58
13.59

0.50
0.40
1.70
0.90
1.09
0.35
0.43
1.16
1.06
1.14
1.89
1.21
1.47
1.75
0.46
0.33
0.33
1.03
0.54
2.81
2.84
0.35
0.65
0.72
0.79
1.02
0.29
2.96
3.29
0.59
1.26
1.71
0.40
0.97
2.16
0.35
0.66
2.18
0.96
0.64
0.66
1.96

56.16
30.45
95.67
133.16
80.45
39.88
53.54
53.37
29.03
75.19
199.25
58.73
88.49
69.03
43.77
86.16
39.92
81.65
46.25
108.54
180.49
35.38
114.88
53.55
29.49
99.92
24.65
144.86
198.88
39.3
71.64
184.03
41.07
53.65
109.14
32.78
109.13
57.50
194.39
22.65
60.21
74.69

4.72
3.51
6.30
3.70
3.40
2.56
2.63
1.57
1.44
3.16
5.90
2.27
3.29
1.86
4.23
4.23
2.69
2.32
2.86
4.47
5.33
3.09
3.43
2.46
2.15
2.67
2.94
4.65
1.67
3.63
1.95
6.56
2.28
5.79
4.28
5.73
4.03
2.79
3.67
1.72
3.73
6.70

578
409
725
630
579
681
487
585
369
540
639
468
498
559
468
1012
520
540
455
609
622
589
798
571
441
876
440
535
622
423
605
571
576
506
696
441
1068
431
559
359
628
550

49
47
48
18
24
44
24
17
18
23
19
18
19
15
45
50
35
15
28
25
18
51
24
26
32
23
53
17
5
39
16
20
32
55
27
77
39
21
11
27
39
49

0.009
0.013
0.017
0.007
0.013
0.009
0.008
0.021
0.035
0.015
0.009
0.020
0.016
0.025
0.010
0.004
0.008
0.012
0.012
0.025
0.015
0.010
0.006
0.013
0.026
0.010
0.012
0.020
0.016
0.015
0.017
0.009
0.010
0.018
0.019
0.011
0.006
0.037
0.005
0.027
0.011
0.026

439
438
441
440
441
440
441
441
437
437
447
438
441
439
440
440
438
441
442
440
445
440
441
443
439
443
441
440
454
437
438
454
436
434
435
436
441
433
453
436
442
443

a
b
c
d
e
f
g
h
i

TOC = total organic carbon.


S1 = free hydrocarbons.
S2 = pyrolysable hydrocarbons.
S3 = carbon dioxide.
HI = hydrogen index.
OI = oxygen index.
PI = productivity index.
Tmax = temperature of maximum S2.
HC = hydrocarbon.

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

Fig. 3. Plot of TOC wt.% versus S2 mg HC/g rock indicating the kerogen types.

(NASC; Gromet et al., 1984); while, Al2O3, MgO, TiO2 and K2O contents
are lower than both of them; MnO contents of oil shale samples vary
from 0.09% to 0.29%, and the average being 2.57%, close to average
shale (0.36%), but much lower than that of NASC (2.98%).
Particularly, Dahuangshan oil shale is obviously enriched in P,
with a mean value of 0.71%, which is much higher than that of average shale (0.19%) and NASC (0.13%). This phenomenon is probably
caused by the input of the volcanic ash which is indicated by a tuffaceous composition of the interbedded mudstone and carbonate layers
(Fig. 5). Volcanic ash provided ample nutrients for growth of algae
leading to formation of a reducing depositional environment favorable
for preservation of organic matter. As a result, the oil yield values of
the Dahuangshan oil shale samples are relatively high, with an average
of 15.0%, which is even higher than the mean value of the Green River
oil shales (11.44%, Ruhl, 1982). The analysis results indicate that SiO2
is the dominant constituent of the outcrop samples, the variation of
which inuences directly on the contents of other elements (Table 3);
therefore, due to a dilution effect SiO2 exhibits a negative correlation
with most major elements and trace elements, which is so-called
Diluent Effect of SiO2.
Al2O3 is also an important component in surface sediments, the
contents of Al2O3 and SiO2 always change in the reverse direction.

45

Meanwhile, SiO2 mainly occurs in compositions of coarse sediments,


whereas Al2O3 is the characteristic component of clay minerals,
therefore, there is a signicant negative correlation between them
(r = 0.951). The elements K and Ti are also mainly associated
with clay minerals, therefore, there is a highly signicant correlation
between Al2O3 and K2O (r = 0.959), Al2O3 and TiO2 (r = 0.928),
K2O and TiO2 (r = 0.812). On the contrary, the concentration of SiO2
correlates negatively with TiO2 (r = 0.87) and K2O (r = 0.954).
Table 3 also shows that there is a signicant correlation between Al2O3 and Fe2O3 (r = 0.847), and MgO (r = 0.836), and
MnO (r = 0.819), and Cu (r = 0.946), and Ba (r = 0.787), and Co
(r = 0.952), and Ni (r = 0.905). The correlation coefcients exceed
0.8 mostly, which reects the fact that these elements are associated
with clay minerals. Besides, the correlation between Fe2O3 and MnO,
Co, Ni is also strong (r = 0.858, 0.903, 0.786, respectively), which
reects the similarity on geochemical behavior of these iron group
elements.
4.4. REE distribution patterns
The concentration of REEs and some other trace elements of 10
samples from the Dahuangshan oil shale are presented in Table 4.
The content of total rare earth elements (REE) varies considerably,
ranging from 31.49 to 111.18 g/g. The weighted mean value is
79.48 g/g, which is higher than that of the average REE contents of US
coals (53.59 g/g; Finkelman, 1993) and the marine oil shale from the
Changshe Mountain area, northern Tibet, China (68.19 g/g; Fu et al.,
2010), but lower than those of world-wide black shales (134.19 g/g;
Ketris and Yudovich, 2009), common Chinese coals (162.51 g/g; Dai
et al., 2008), and the NASC (167.41 g/g; Haskin et al., 1968).
The concentration of the light rare earth elements (LREEs) is higher
than that of the heavy rare earth elements (HREEs), which is in accordance with the general distribution of REEs in shales (e.g. Condie,
1991; Fu et al., 2010; Gromet et al., 1984; Ketris and Yudovich, 2009).
The LREE/HREE ratios range from 5.16 to 9.22. The value of (La/Yb)N
has the same signicance with the LREE/HREE ratios, ranging from
6.67 to 10.69 (Table 4), indicating the enrichment of LREEs. (La/Sm)N
values vary from 2.70 to 5.95, showing a strong fractional degree
among LREEs, suggesting the deep water sedimentary environment of
oil shale.
All oil shale samples exhibit a negative Eu anomaly (Table 4), with
a mean Eu value of 0.64. The Ce values of all samples vary from 0.81
to 1.11, showing a slightly or negligibly negative anomaly in some
plies.
The oil shale samples exhibit two slightly different types A and B
of the chondrite-normalized REE patterns although they show coherent, subparallel REE patterns (Fig. 6). Both types A and B are characterized by distinctly sloping LREE trends (LaN/SmN = 2.705.95)
accompanied by at HREE trends, and there is no notable difference
in Eu between types A and B oil shales. However, No Ce depletion
is seen in the REE distribution patterns of type A (samples O-11,
O-13, O-20, O-28, O-29, and O-39), showing negligible Ce anomalies.
A small to medium Eu depletion is observed and shows a small or
middle V shape in the REE distribution patterns. In contrast, Type
B (samples O-04, O-21, O-32, and O-35) shows a little negative Ce
anomaly with slight Ce depletion (Fig. 6B). At the same time, Type B
exhibits obviously Nd anomaly with a medium increase and shows a reversed V shape. The average values of REE, LREE/HREE, (La/Yb)N, and
(La/Sm) N of type A samples are relatively higher than those of type B
samples.
4.5. Molecular geochemistry of organic matter

Fig. 4. Plot of Tmax C versus HI mg HC/g TOC identifying the kerogen types.

4.5.1. n-Alkanes
Aliphatic hydrocarbons are dominated by n-alkanes. Gas chromatograms of saturated hydrocarbons from Dahuangshan oil shales

46

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

Table 2
Concentrations of ash, oil yield, total sulfur, organic sulfur and major elements in samples from the Dahuangshan oil shale (unit in %).
Sample no.

Oil yield

Ad

St,d

So,d

SiO2

Fe2O3

Al2O3

CaO

MgO

K2O

Na2O

TiO2

P2O5

MnO

O-04
O-11
O-13
O-20
O-21
O-28
O-29
O-32
O-35
O-39
Ave.
Ave. shalea
NASCb

7.8
4.9
9.9
12.2
20.5
26.6
15.1
17.5
9.4
26.2
15.0

52.35
64.33
55.08
67.14
65.45
62.37
71.03
64.18
56.91
50.2
60.90

0.88
0.48
0.89
0.59
0.32
0.35
0.31
0.34
0.52
0.58
0.53

0.31
0.21
0.3
0.15
0.13
0.09
0.09
0.18
0.23
0.23
0.19

66.14
69.01
63.84
64.7
71.09
71.5
76.78
70.14
65.42
67.29
68.59
64.21
64.8

6.68
4.96
6.84
5.34
5.18
4.32
4.76
4.76
5.98
5.46
5.43
6.71
5.66

12.39
10.26
12.88
11.3
8.6
7.82
6.37
9.4
12.69
10.12
10.18
17.02
16.9

1.33
3.14
3.8
5.02
3.41
4.1
3.3
3.7
3.6
3.96
3.54
3.44
3.63

2.34
1.97
2.08
2.09
1.84
1.81
1.73
1.86
2.63
1.98
2.03
2.7
2.86

2.52
2.34
2.82
2.76
2.05
1.82
1.5
2.14
2.75
2.12
2.28
3.58
3.97

2.2
0.7
1.93
2.28
0.61
0.56
0.44
0.73
1.46
1.23
1.21
1.44
1.14

0.68
0.58
0.64
0.55
0.51
0.48
0.41
0.5
0.62
0.6
0.56
0.72
0.70

0.86
1.15
0.98
0.3
0.42
0.48
0.36
0.78
0.86
0.92
0.71
0.19
0.13

0.22
0.14
0.29
0.13
0.18
0.11
0.09
0.18
0.25
0.20
0.18
0.5
0.06

Ad, ash yield, dry basis; St,d, total sulfur, dry basis; So,d, organic sulfur, dry basis.
a
From Clarke (1924).
b
From Gromet et al. (1984).

are presented in Fig. 7. The n-alkanes patterns of oil shales from


Dahuangshan area are dominated by short (nC15nC19, Bechtel
et al., 2002)- to middle (nC21nC25)-chain n-alkanes with highest
relative intensities in the nC20 to nC22 range without a marked carbon
preference index (CPI b 1.3) in the nC23 to nC31 range. The results are
different from the long-chain n-alkanes with a marked odd over
even preference in Tertiary oil shales in the central Tibetan plateau
(Wang et al., 2011). Brassell et al. (1978) reported that the single
peak with the maximum peak carbon of nC18nC24, low CPI value, indicated the organisms were derived from phytoplankton, zooplankton, benthic bacteria with no photosynthesis and terrestrial plants.
In addition, Allen et al. (1971) indicated that microbial reworked
organic matter in sediments is characterized by low CPI values at
low maturation stage. Therefore, the organic matter in Dahuangshan
oil shales mainly comes from thallogen microorganisms living in a

deep water environment with bottom waters of elevated salinity


(Zhu et al., 2005). The values Paq of Ficken et al. (2000), where
Paq = (C23 + C25) / (C23 + C25 + C29 + C31), obtained from 10
samples (averaging 0.75) are close to the data measured in submerged/oating macrophytes (Ficken et al., 2000). The pristane/
phytane (Pr/Ph) ratio is a commonly-used parameter for the study
of oxic/anoxic conditions and sources of organic matter (Didyk
et al., 1978; Escobar et al., 2011), although some studies have demonstrated that there are multiple possible relationships between
depositional environment and Pr/Ph ratio (ten Haven et al., 1987).
The pristane/phytane ratios in Dahuangshan oil shales vary from
0.41 to 0.91 and average 0.60. According to Peters and Moldowan
(1993), Pr/Ph b 0.6 within the oil-generative window indicates
anoxic conditions; whereas Pr/Ph > 3 indicates suboxic to oxic depositional environments. In Dahuangshan area, Pr/Ph ratios in most

Td

Do

Td

Td
Q

Ta
Sd
Do

200 m

Td

Ta

200m

Td

Td
Q

Td
Qa
Td

200m

200m

Ta

Fig. 5. Minerals in the Dahuangshan prole samples (existing as thin bed or thin interbedded with oil shale), Thin section, Polarized light. A limestone with tuff debris (Td), tuffaceous matrix (Ta), and quartz (Q); B dolomite with tuff debris, dolomite (Do), quartz, and siliceous debris (Sd); C tuffaceous mudstone with tuff debris and tuffaceous matrix;
D tufte with tuff debris, tuffaceous matrix, quartz, and authigenic quartz (Qa).

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

47

Table 3
Correlation coefcient values of major elements and trace elements.

SiO2
Al2O3
Fe2O3
CaO
MgO
Na2O
K2O
TiO2
P2O5
MnO
Ba
Sr
V
Cu
Co
Ni

SiO2

Al2O3

Fe2O3

CaO

MgO

Na2O

K2O

TiO2

P2O5

MnO

Ba

Sr

Cu

Co

Ni

1.000
0.951
0.735
0.087
0.715
0.850
0.954
0.870
0.455
0.730
0.867
0.412
0.287
0.898
0.872
0.843

1.000
0.847
0.150
0.836
0.836
0.959
0.928
0.564
0.819
0.787
0.333
0.041
0.946
0.952
0.905

1.000
0.402
0.690
0.792
0.739
0.854
0.451
0.858
0.429
0.032
0.154
0.881
0.903
0.786

1.000
0.260
0.061
0.053
0.372
0.378
0.257
0.353
0.469
0.838
0.132
0.228
0.132

1.000
0.660
0.762
0.767
0.395
0.656
0.697
0.478
0.001
0.876
0.904
0.739

1.000
0.830
0.751
0.148
0.547
0.696
0.130
0.148
0.852
0.835
0.633

1.000
0.812
0.415
0.703
0.892
0.390
0.147
0.874
0.867
0.825

1.000
0.687
0.809
0.600
0.146
0.074
0.896
0.900
0.900

1.000
0.619
0.179
0.013
0.297
0.452
0.465
0.773

1.000
0.431
0.301
0.008
0.803
0.843
0.816

1.000
0.601
0.422
0.737
0.696
0.638

1.000
0.609
0.320
0.330
0.228

1.000
0.143
0.050
0.051

1.000
0.988
0.881

1.000
0.860

1.000

of the samples are b0.6, indicating an anoxic deposition. In addition,


syngenetic pyrite can be found in studied samples, which can also
prove the anoxic sedimentary environment of these samples (e.g. Dai
et al., 2002; Fu et al., 2010; Hackley et al., 2009; Kara-Glbay et al.,
2012; Nowak, 2007; Sabel et al., 2005).

hopanes above C31 gradually decreases. Homohopanes (C31C35) are


believed to be derived from bacteriohopanetetrol as well as from
other hopanoids in bacteria (Ourisson et al., 1984). The homohop indices [C35 homohopane/(C31C35) homohopanes] of all samples are more
than 0.06, indicating a reducing depositional environment (Peters and
Moldowan, 1991).
Gammacerane occurs in small amounts for all samples (Fig. 8). A
high gammacerane index [Gammacerane/C30-hopane] is interpreted
to indicate highly reducing, hypersaline conditions during deposition.
However, relatively high gammacerane abundances are also seen in
freshwater lacustrine sediments, and Sinninghe Damst et al. (1995)
thus proposed that gammacerane index is in fact an indicator for
water column stratication. The compound is abundant in saline lacustrine deposits just because the water columns in hypersaline depositional environments are often density stratied (Zhu et al., 2005). The
gammacerane index varies from 0.08 to 0.36 in the Dahuangshan oil
shale, combining with the low Pr/Ph ratio and other biomarker parameters, thus probably indicates a moderate saline lacustrine environment.

4.5.2. Terpanes
The distribution and relative abundances of pentacyclic and tricyclic
terpanes obtained from m/z 191 ion chromatograms are shown in Fig. 8
and their parameters are given in Table 5. Triterpenoids are relatively rare
in Dahuangshan oil shale samples. Traces of unsaturated triterpenes
are found in most of the samples. The hopanoids are dominated by the
presence of C30-hopane, C29-norhpane, 17(H)-trisnorhopane (Tm),
and a considerable quantity of homohopanes (C31C35) (Fig. 8).
Hopanes are ubiquitous constituents of sedimentary organic matter
(Piedad-Snchez et al., 2004; Zumberge, 1987). They are derived from a
degraded bacteriohopane C35 (Ourisson et al., 1979). As shown in Fig. 8,
the hopane C30 is the highest whereas the relative abundance of
Table 4
Rare earth element contents (in g/g) in samples and associated geochemical parameters.
Element

O-4

O-11

O-13

O-20

O-21

O-28

O-29

O-32

O-35

O-39

La
Ce
Pr
Nd
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu
Y
V
Ni
Sr
Ba
Cu
Co
REE
L/H
(La/Yb)N
Ce
Eu
(La/Sm)N

18.1
33.8
4.80
19.9
4.21
0.84
4.22
0.68
3.68
0.79
2.12
0.32
2
0.31
20.7
65.7
41.2
129
221
68.20
13.41
95.77
5.78
6.10
0.87
0.61
2.70

6.27
11.7
1.47
5.67
1.16
0.25
1.30
0.20
1.19
0.26
0.85
0.13
0.88
0.16
8.25
101
36.6
110
97.2
48.67
8.58
31.49
5.34
4.80
0.93
0.62
3.40

21.9
42.8
5.31
21.2
4.40
0.92
4.33
0.69
3.87
0.77
2.30
0.33
2.04
0.32
21.1
128
43.5
146
268
69.81
13.28
111.18
6.59
7.24
0.96
0.64
3.13

16.5
36.6
4.03
16.8
3.35
0.74
3.53
0.58
3.11
0.67
1.79
0.28
1.64
0.27
17.2
184
39.7
185
340
63.25
11.57
89.89
6.57
6.78
1.08
0.66
3.10

23.0
35.5
4.79
26.2
2.91
0.63
3.44
0.52
3.03
0.63
1.95
0.28
1.82
0.28
17.9
133
37.8
163
195
52.87
9.74
104.98
7.78
8.52
0.81
0.61
4.97

13.4
25.4
3.17
12.7
2.60
0.56
2.95
0.46
2.57
0.56
1.73
0.26
1.69
0.26
16.4
143
37.2
170
212
52.69
9.23
68.31
5.52
5.35
0.94
0.62
3.24

9.68
17.4
2.24
8.8
1.83
0.44
1.86
0.30
1.70
0.37
1.08
0.17
1.05
0.16
9.59
86.3
41.5
133
235
55.76
10.05
47.08
6.04
6.22
0.90
0.73
3.33

17.6
25.3
2.85
14.7
1.86
0.41
1.89
0.29
1.71
0.36
1.10
0.17
1.11
0.17
10
120
38.4
198
191
52.10
9.82
69.52
9.22
10.69
0.86
0.67
5.95

17.6
34.5
4.56
18.3
3.72
0.84
4.08
0.65
3.90
0.88
2.61
0.38
2.51
0.39
24.7
147
42.9
237
322
71.24
13.89
94.92
5.16
4.73
0.93
0.66
2.98

12.6
28.3
2.98
11.2
2.30
0.46
2.36
0.35
1.99
0.42
1.30
0.19
1.35
0.22
11.9
189
41.5
176
207
62.08
11.28
66.02
7.07
6.29
1.11
0.60
3.45

L/H = LREE/HREE; (La/Yb)N, (La/Sm)N; subscripts N stands for chondrite-normalized value (Boynton, 1984); Ce/Ce* = CeN/(LaN PrN)0.5; Eu/Eu* = EuN/(SmN GdN)0.5; subscripts N
stands for chondrite-normalized value (Boynton, 1984).

48

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

Fig. 6. Distribution patterns of rare earth elements in oil shale samples by chondrite-nomalized. REE patterns.

In general, Pr/nC17 and Ph/nC18 ratios would decrease along with the
increasing in maturity of organic matter (Moldowan et al., 1985). The
Pr/nC17 ratios range between 0.31 and 1.04, and Ph/nC18 ratios vary between 0.45 and 1.45 (Table 5), which are lower than those of immature
oil shale in the central Tibetan plateau (Wang et al., 2011), showing the
early maturation stage of oil shale in Dahuangshan area. This result is
consistent with the relatively high Tmax values as discussed above.
4.5.3. Steranes
The steranes composition could be correlated with the type of
environment (Huang and Meinschein, 1979). They proposed that a
dominance of C27 sterols (steranes) mainly derive from algae, while
the C29 sterols are more typically associated with land plants. Volkman
(1986) indicated that the low C28 levels are typical of limnic environments. However, they also considered that microalgae or cyanobacteria
can also be important sources of C29 sterols. From the m/z 217 mass chromatograms of our samples, the relative abundances of the C27, C28, and
C29 steranes and their 20S and 20R epimers have been determined
(Fig. 8; Table 5). Dahuangshan oil shale samples show a higher proportion of C27 (2455%) compared to C29 (2647%) and C28 (1329%)

steranes, except sample O-04 having C29 > C27 steranes distribution,
reecting a high contribution of aquatic algae (Peters and Moldowan,
1993), which is consistent with the n-alkanes distribution. The predominance of C29 steranes in sample O-04 may reect a high contribution of
bacteria or/and microorganisms (Riboulleau et al., 2007; Vandenbroucke
and Bchar, 1988).
Moreover, 20(S)/(20S + 20R) and /( + ) sterane ratios
increase with increasing maturity of organic matter (Hunt, 1996;
Goodarzi et al., 1989; Seifert and Moldowan, 1980). As shown in
Table 5, 20(S)/(20S + 20R) and /( + ) ratios are in the range
0.20.34, and 0.150.32, respectively, indicating the early maturation
stage of organic matter (Grantham, 1986; Goodarzi et al., 1989).
5. Conclusions
(1) Two types of facies sequences are developed in section No.2.
from Dahuangshan area. The middleupper part of the section
consists of dark-black oil shale interbedded by thin-bedded argillaceous dolomite, silty claystone, tuff, limestone, and dolomitic
marl, containing a spot of plant stem fossils, and an abundance

Fig. 7. Mass chromatograms (m/z = 85) for the saturated alkanes of selected samples from the Dahuangshan oil shale.

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

49

Fig. 8. m/z 191 and m/z 217 ion fragmentograms of the saturated fractions showing the distribution of the terpanes and steranes for selected oil shale samples.

of shtail and sh skeleton, while, the lower part is lack of oil


shale and fossils. The sedimentary characteristics of oil shale suggest a lacustrine deposition.
(2) Dahuangshan oil shales contain high values of TOC, S2, HI, and
oil yield, indicating the oil shales have excellent source rock
potential. Kerogen types of organic matter are type II and I,
which further verify the oil prone source rock potential. Tmax
values, Pr/nC17 and Ph/nC18 ratios, 20(S)/(20S + 20R) and
/( + ) sterane ratios show an early to medium maturation stage of organic matter.
(3) The oil shale samples are rich in SiO2 (68.59%), Al2O3 (10.18%), and
Fe2O3 (5.43%). Element P is especially enriched compared with average shale and North American Shale Composite (NASC), which is
probably caused by the input of the volcanic ash, representing an
excellent situation for organic matter production and enrichment.

(4) The content of REE in oil shale samples vary considerably,


ranging from 31.49 to 111.18 g/g, with a weighted mean
value of 79.48 g/g, which are higher than those of US coals,
but lower than those of world-wide black shales, common
Chinese coals and the NASC.
(5) The contents of Al2O3 and SiO2 always change in the reverse
direction. There is a highly signicant correlation between
Al2O3 and K2O (r = 0.959), Al2O3 and TiO2 (r = 0.928), and
K2O and TiO2 (r = 0.812). On the contrary, the concentration
of SiO2 correlates negatively with TiO2 (r = 0.87) and K2O
(r = 0.954). Moreover, the signicant correlations between
Al2O3 and Fe2O3, MgO, MnO, Cu, Ba, Co, and Ni are because
of their strong afnity to the clay minerals, and the signicant
correlations between Fe2O3 and MnO, Co, and Ni are due to
their similar geochemical behavior.

50

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151

Table 5
Organic geochemical data for extracts of samples from the Dahuangshan oil shale.
Sample no.

CPI

Pr/Pha

Pr/nC17

Ph/nC18

Gammacerane indexb

Homohop indexc

20(S)/(20S + 20R)d

/( + )

O-04
O-11
O-13
O-20
O-21
O-28
O-29
O-32
O-35
O-39

1.12
0.98
1.24
1.08
0.93
1.01
1.05
0.99
1.16
1.01

0.41
0.59
0.82
0.60
0.91
0.56
0.57
0.61
0.42
0.52

0.97
0.59
0.45
0.31
0.91
0.56
0.37
0.72
1.04
0.82

1.24
1.02
0.98
0.47
0.56
1.02
0.45
0.67
1.45
0.85

0.09
0.23
0.26
0.18
0.36
0.11
0.13
0.08
0.12
0.15

0.112
0.066
0.061
0.077
0.084
0.106
0.109
0.990
n.d.
0.145

0.21
0.23
0.28
0.24
0.34
0.31
0.21
0.29
0.20
0.22

0.19
0.24
0.32
0.18
0.28
0.20
0.15
0.26
0.17
0.18

%C27f

%C28g

%C29h

24
55
37
50
48
44
53
47
46
52

29
14
27
24
16
25
13
22
22
20

47
31
36
26
36
31
34
31
32
28

n.d. not detected.


a
Pr/Ph = pristane/phytane ratio.
b
Gammacerane index = Gammacerane/C30-hopane.
c
Homohop index = C35, (22S + 22R)-/C31C35, (22S + 22R)-hopanes.
d
20(S)/(20S + 20R) = C29,20S/C29 , (20S + 20R) steranes.
e
/( + ) = C29-regular sterane (20R + 20S)/(20S + 20R + 20R + 20S) isomer ratio.
f
%C27 = %C27/C27C29 steranes.
g
%C28 = %C28/C27C29 steranes.
h
%C2 9 = %C29/C27C29 steranes.

(6) All selected oil shale samples are characterized by distinctly


sloping LREE trends (LaN/SmN = 2.705.95) accompanied by at
HREE trends, with distinct Eu negative anomalies (0.600.73).
Two slightly different patterns of REEs in the oil shale samples
are distinguished by the difference in Ce depletion and Nd anomaly. Compared to type B, the average values of REE, LREE/HREE,
(La/Yb)N, and (La/Sm)N of type A samples are relatively higher.
(7) Dahuangshan oil shale samples are characterized by short- to
middle-chain n-alkanes, low CPI values (0.931.24), single peak
composed of nC20 or nC22, low Pr/Ph (0.410.91), relatively high
Homohop index (0.0610.99), and high concentrations of C27
sterane, which is consistent with a reducing, paleo-lake environment of moderate salinity with a prevalent contribution of algae
and microorganisms to organic matter accumulation.
Acknowledgments
This work was subsidized by the Key Project of the National Science &
Technology (2011ZX05034-001), the National Basic Research Program
of China (973) (2009CB219604), the Postdoctoral Science Fund of
China (2011 M500433) and the Scientic Research Foundation of Key
Laboratory of Coalbed Methane Resources and Reservoir Formation Process, Ministry of Education (China University of Mining and Technology)
(2012-001).
References
Alaug, A.S., 2011. Source rocks evaluation, hydrocarbon generation and palynofacies
study of Late Cretaceous succession at 16/G-1 offshore well in Qamar Basin, eastern
Yemen. Arabian Journal of Geosciences 4, 551566.
Allen, J.E., Fornery, F.W., Markovetz, A.J., 1971. Microbial degradation of n-alkanes.
Lipids 6, 448452.
Amijaya, H., Schwarzbauer, J., Littke, R., 2006. Organic geochemistry of the Lower
Suban coal seam, South Sumatra Basin, Indonesia: palaeoecological and thermal
metamorphism implications. Organic Geochemistry 37, 261279.
Bechtel, A., Sachsenhofer, R.F., Kolcon, I., Kolcon, R., Otto, A., Pttmann, W., 2002. Organic geochemistry of the Lower Miocene Oberdorf lignite (Styrian Basin, Austria):
its relation to petrography, palynology and the palaeoenvironment. International
Journal of Coal Geology 51, 3157.
Bolou-Bi, E.B., Poszwa, A., Leyval, C., Vigier, N., 2010. Experimental determination of magnesium isotope fractionation during higher plant growth. Geochimica et Cosmochimica
Acta 74, 25232537.
Boynton, W.V., 1984. Geochemistry of the rare earth elements: meteorite studies. In:
Henderson, P. (Ed.), Rare Earth Element Geochemistry. Elsevier, pp. 63114.
Brassell, S.C., Eglinton, G., Maxwell, J.R., Philp, R.P., 1978. Natural Background of Alkanes in
the Aquatic Environment. In: Hutzinger, L.H., van Lelyveld, O., Zoeteman, B.C.J. (Eds.),
Pergamon, Oxford, pp. 6986.

Carroll, A.R., 1998. Upper Permian lacustrine organic facies evolution, Southern Junggar
Basin, NW China. Organic Geochemistry 28, 649667.
Carroll, A.R., Brassell, S.C., Graham, S.A., 1992. Upper Permian Lacustrine oil shales,
southern Junggar Basin, northwest China. American Association of Petroleum Geologists Bulletin 76, 18741902.
Clarke, F.W., 1924. The data of geochemistry. U.S. Geological Survey Bulletin 770.
Clayton, J.L., King, J.D., 1987. Effects of weathering on biological marker and aromatic
hydrocarbon composition of organic matter in Phosphoria shale outcrop. Geochimica
et Cosmochimica Acta 51, 21532157.
Clayton, J.L., Ryder, R.T., 1984. Organic geochemistry of black shales and oils in the
Minnelusa Formation (Permian and Pennsylvanian), Powder River Basin, Wyoming.
In: Woodward, J., Meissner, F.F., Clayton, J.L. (Eds.), Hydrocarbon Source Rocks of
the Greater Rocky Mountain Region. Rocky Mt. Assoc. Geol, Denver, CO, pp. 231253.
Condie, K.C., 1991. Another look at rare earth elements in shales. Geochimica et
Cosmochimica Acta 55, 25272531.
Cooper, B.S., Barnard, P.C., 1984. Source rocks and oils of the central and northern North
Sea. In: Demaison, G., Murris, R.J. (Eds.), Petroleum Geochemistry and Basin Evaluation, AAPG Memoir 35. American Association of Petroleum Geologists, Tulsa, pp. 114.
Cornford, C., Gardner, P., Burgess, C., 1998. Geochemical truths in large data sets: I. Geochemical screening data. Organic Geochemistry 29, 519530.
Dahl, B., Bojesen-Koefoed, J., Holm, A., Justwan, H., Rasmussen, E., Thomsen, E., 2004. A
new approach to interpreting Rock-Eval S2 and TOC data for kerogen quality assessment. Organic Geochemistry 35, 14611477.
Dai, S., Ren, D., Tang, Y., Shao, L., Li, S., 2002. Distribution, isotopic variation and origin
of sulfur in coals in the Wuda coaleld, Inner Mongolia, China. International Journal of Coal Geology 51, 237250.
Dai, S.F., Li, D., Chou, C.L., Zhao, L., Zhang, Y., Ren, D.Y., Ma, Y.W., Sun, Y.Y., 2008. Mineralogy
and geochemistry of boehmite-rich coals: new insights from the Haerwusu Surface
Mine, Junggar Coaleld, Inner Mongolia, China. International Journal of Coal Geology
74, 185202.
Demaison, G.J., Moore, G.T., 1980. Anoxic environments and oil source bed genesis.
AAPG Bulletin 64, 11791209.
Didyk, B.M., Simoneit, B.R.T., Brassell, S.C., Eglinton, G., 1978. Organic geochemical indicators of paleoenvironmental conditions of sedimentation. Nature 272, 216222.
Escobar, M., Mrquez, G., Inciarte, S., Rojas, J., Esteves, I., Malandrino, G., 2011. The
organic geochemistry of oil seeps from the Sierra de Perij eastern foothills, Lake
Maracaibo Basin, Venezuela. Organic Geochemistry 42, 727738.
Espitali, J., Deroo, G., Marquis, F., 1985. La pyrolyse Rock-Eval et ses applications
(deuximepartie). Revue Institut Francais du Ptrole 40, 755784.
Ficken, K.J., Li, B., Swain, D.L., Eglinton, G., 2000. An n-alkane proxy for the sedimentary
input of submerged/oating freshwater aquatic macrophytes. Organic Geochemistry
31, 745749.
Finkelman, R.B., 1993. Trace and minor elements in coal. In: Engel, M.H., Macko, S.A.
(Eds.), Organic Geochemistry. Plenum, New York, NY, pp. 593607.
Fu, X.G., Wang, J., Zeng, Y.H., Tan, F.W., Feng, X.L., 2010. REE geochemistry of marine oil
shale from the Changshe Mountain area, northern Tibet, China. International Journal
of Coal Geology 81, 191199.
Goodarzi, F., Brooks, P.W., Embry, A.F., 1989. Regional maturity as determined by
organic petrography and geochemistry of the Schei Point Group (Triassic) in the
western Sverdrup Basin, Canadian Arctic Archipelago. Marine and Petroleum Geology
6, 290302.
Graham, S.A., Brassell, S., Carroll, A.R., Xiao, X., Demaison, G., Mcknight, C.L., Liang, Y.,
Chu, J., Hendrix, M.S., 1990. Characteristics of selected petroleum source rocks,
Xinjiang Uygur Autonomous Region, northwest China. AAPG Bulletin 74, 493512.
Grantham, P.J., 1986. Sterane isomerisation and moretane/hopane ratios in crude oils
derived from Tertiary source rocks. Organic Geochemistry 9, 293304.

S. Tao et al. / International Journal of Coal Geology 115 (2013) 4151


Gromet, L.P., Dymek, R.F., Haskin, L.A., Korotev, R.V., 1984. The North American shale
composite: its compilation, major and trace element characteristics. Geochimica
et Cosmochimica Acta 48, 24692482.
Hackley, P.C., Guevara, E.H., Hentz, T.F., Hook, R.W., 2009. Thermal maturity and organic
composition of Pennsylvanian coals and carbonaceous shales, north-central Texas:
implications for coalbed gas potential. International Journal of Coal Geology 77,
294309.
Hamburg, A., 2011. Analysis of energy development perspectives. Oil Shale 28, 367371.
Haskin, L.A., Haskin, M.A., Frey, F.A., Wilderman, T.R., 1968. Relative and absolute
terrestrial abundances of the rare earths. In: Ahrens, L.H. (Ed.), Origin and Distribution of the Elements. Pergamon, Oxford, pp. 889912.
Huang, W.Y., Meinschein, W.G., 1979. Sterols as ecological indicators. Geochimica et
Cosmochimica Acta 43, 739745.
Hunt, J.M., 1996. Petroleum Geochemistry and Geology, 2nd ed. Freeman, New York 743.
Jiao, Y.Q., Wu, L.Q., He, M.C., Roger, M., Wang, M.F., Xu, Z.C., 2007. Occurrence, thermal
evolution and primary migration processes derived from studies of organic matter
in the Lucaogou source rock at the southern margin of the Junggar Basin, NW
China. Science in China Series D 50, 114123.
Kara-Glbay, R., Krmac, M.Z., Korkmaz, S., 2012. Organic geochemistry and depositional
environment of the Aptian bituminous limestone in the Kale Gmhane area (NETurkey): an example of lacustrine deposits on the platform carbonate sequence. Organic Geochemistry 49, 617.
Ketris, M.P., Yudovich, Y.E., 2009. Estimations of clarkes for carbonaceous biolithes:
world average for trace element contents in black shales and coals. International
Journal of Coal Geology 78, 135148.
Korkmaz, S., Glbay, R.K., 2007. Organic geochemical characteristics and depositional
environments of the Jurassic coals in the eastern Taurus of Southern Turkey. International Journal of Coal Geology 70, 292304.
Langford, F.F., Blanc-Valleron, M., 1990. Interpreting Rock-Eval pyrolysis data using
graphs of pyrolizable hydrocarbons vs. total organic carbon. American Association
of Petroleum Geologists Bulletin 74, 799804.
Leythaeuser, D., 1973. Effects of weathering on organic matter in shales. Geochimica et
Cosmochimica Acta 37, 113120.
Li, S.Y., 2012. The developments of Chinese oil shale activities. Oil Shale 29, 101102.
Littke, R., Klussmann, U., Krooss, B., Leythaeuser, D., 1991. Quantication of loss of
calcite, pyrite, and organic matter due to weathering of Toarcian black shales and
effects on kerogen and bitumen characteristics. Geochimica et Cosmochimica
Acta 55, 33693378.
Littke, R., Jendrzejewski, L., Lokay, P., Wang, S.Q., Rullktter, J., 1998. Organic geochemistry and depositional history of the BarremianAptian boundary interval in the
Lower Saxony Basin, northern Germany. Cretaceous Research 19, 581614.
Liu, Z.J., Dong, Q.S., Zhu, J.W., Guo, W., Ye, S.Q., Liu, R., Zhang, H.L., 2007. Oil shale resources and its distribution in China. Oil and Gas Resources 3, 2628.
Moldowan, J., Seifert, W., Gallegos, E., 1985. Relationship between petroleum composition and depositional environment of petroleum source rocks. AAPG Bulletin 69,
12551268.
Nowak, G.J., 2007. Comparative studies of organic matter petrography of the late
Palaeozoic black shales from Southwestern Poland. International Journal of Coal
Geology 71, 568585.
Ourisson, G., Albrecht, P., Maxwell, J.R., Wheatley, R.E., 1979. The hopanoids.
Palaeochemistry and biochemistry of a group of natural products. Pure and Applied
Chemistry 51, 709729.
Ourisson, G., Albrecht, P., Rohmer, M., 1984. The microbial origin of fossil fuels. Scientic American 251, 4451.
Peters, K.E., Cassa, M.R., 1994. Applied source rock geochemistry. In: Magoon, L.B., Dow,
W.G. (Eds.), The Petroleum System From Source to Trap: American Association
of Petroleum Geologists Memoir, 60, pp. 93120.
Peters, K.E., Moldowan, J.M., 1991. Effects of source, thermal maturity, and biodegradation on the distribution and isomerization of homohopanes in petroleum. Organic
Geochemistry 17, 4761.
Peters, K.E., Moldowan, J.M., 1993. The Biomarker Guide, Interpreting Molecular Fossils
in Petroleum and Ancient Sediment. Prentice Hall, New Jersey 483664.

51

Piedad-Snchez, N., Surez-Ruiz, I., Martnez, L., Izart, A., Elie, M., Keravis, D., 2004.
Organic petrology and geochemistry of the Carboniferous coal seams from the
Central Asturian Coal Basin (NW Spain). International Journal of Coal Geology 57,
211242.
Riboulleau, A., Schnyder, J., Riquier, L., Lefebvre, V., Baudin, F., Deconinck, J.F., 2007.
Environmental change during the Early Cretaceous in the Purbeck-type Durlston
Bay section (Dorset, Southern England): a biomarker approach. Organic Geochemistry
38, 18041823.
Ruhl, W., 1982. Tar (Extra Heavy Oil) Sand and Oil Shales. Ferdinand Enke, Stuttgart
1149.
Ryu, J.S., Jacobson, A.D., Holmden, C., Lundstrom, C., Zhang, Z.F., 2011. The major ion,
44/40Ca, 44/42Ca, and 26/24Mg geochemistry of granite weathering at pH = 1
and T = 25 C: power-law processes and the relative reactivity of minerals. Geochimica
et Cosmochimica Acta 75, 60046026.
Sabel, M., Bechtel, A.P., ttmann, W., Hoernes, S., 2005. Palaoenvironment of the Eocene
Eckfeld Maar lake (Germany): implications from geochemical analysis of the oil
shales sequence. Organic Geochemistry 36, 873891.
Seifert, W.K., Moldowan, J.M., 1980. The effect of thermal stress on source-rock quality
as measured by hopane stereochemistry. Physics and Chemistry of the Earth 12,
229237.
Sinninghe Damst, J.S., van Duin, A.C.T., Hollander, D., Kohnen, M.E.L., Leeuw, J.W.,
1995. Early diagenesis of bacteriohopanepolyol derivatives: formation of fossil
homohopanoids. Geochimica et Comsmochimica Acta 59, 51415147.
Tao, S., Tang, D.Z., Li, J.J., Xu, H., Li, S., Chen, X.Z., 2010. Indexes in evaluating the grade of
Bogda Mountain oil shale in China. Oil Shale 27, 179189.
Tao, S., Tang, D.Z., Xu, H., Cai, J.L., Gou, M.F., 2011. Retorting process property of oil shale
in the northern foot of Bogda Mountain, China. Oil shale 28, 1928.
Tao, S., Wang, Y.B., Tang, D.Z., Wu, D.M., Xu, H., He, W., 2012a. Organic petrology of
Fukang Permian Lucaogou Formation oil shales at the northern foot of Bogda
Mountain, Junggar Basin, China. International Journal of Coal Geology 99, 2734.
Tao, S., Wang, Y.B., Tang, D.Z., Xu, H., Zhang, B., Deng, C.M., He, W., 2012b. Estimation of
the mineable oil shale amount in West Fukang at the northern foot of Bogda Mountain, Zhunggar Basin, China. Energy Sources, Part A: Recovery, Utilization, and
Environmental Effects 34, 17911800.
Tao, S., Wang, Y.B., Tang, D.Z., Xu, H., Zhang, B., He, W., Liu, C., 2012c. Composition of
the organic constituents of Dahuangshan oil shale at the northern foot of Bogda
Mountain, China. Oil shale 29, 115127.
ten Haven, H.L., de Leeuw, J.W., Rullktter, J., Sinninghe Damst, J.S., 1987. Restricted
utility of the pristane/phytane ratio as palaeoenvironmental indicator. Nature 330,
641643.
Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence. Springer-Verlag,
Berlin, New York 1699.
Vandenbroucke, M., Bchar, F., 1988. Geochemical characterization of the organic matter
from recent sediments by a pyrolysis technique. In: Kelts, K., Fleet, A., Talbot, M.
(Eds.), Lacustrine Petroleum Source Rocks, vol. 40. Geological Society Special Publication, pp. 91101.
Volkman, J.K., 1986. A review of sterol markers for marine and terrigenous organic
matter. Organic Geochemistry 9, 8399.
Wang, L.C., Wang, C.S., Li, Y.L., Zhu, L.D., Wei, Y.S., 2011. Sedimentary and organic geochemical investigation of tertiary lacustrine oil shale in the central Tibetan Plateau:
palaeolimnological and palaeoclimatic signicances. International Journal of Coal
Geology 86, 254265.
Watson, M.P., Hayward, A.B., Parkinson, D.N., Zhang, Z.M., 1987. Plate tectonic history,
basin development and petroleum source rocks deposition onshore China. Marine
and Petroleum Geology 4, 205225.
World Energy Council, 2007. 2007 Survey of Energy Resources. World Energy Council,
London.
Zhu, Y.M., Weng, H.X., Su, A.G., Liang, D.G., Peng, D.H., 2005. Geochemical characteristics of Tertiary saline lacustrine oils in the Western Qaidam Basin, northwest China.
Applied Geochemistry 20, 8751889.
Zumberge, J.E., 1987. Terpenoid biomarker distributions in low maturity crude oils.
Organic Geochemistry 11, 479496.

You might also like