Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Organic Geochemistry 42 (2011) 323339

Contents lists available at ScienceDirect

Organic Geochemistry
journal homepage: www.elsevier.com/locate/orggeochem

Lacustrine source rock deposition in response to co-evolution


of environments and organisms controlled by tectonic subsidence
and climate, Bohai Bay Basin, China
Fang Hao a,b,, Xinhuai Zhou c, Yangming Zhu d, Yuanyuan Yang a
a

State Key Laboratory of Petroleum Resources and Prospecting, China University of Petroleum, Fuxue Road No. 18, Changping, Beijing 102249, China
Key Laboratory of Tectonics and Petroleum Resources, Ministry of Education, China University of Geosciences, Wuhan 430074, China
Department of Earth Sciences, Zhejiang University, Hangzhou, Zhejiang 310027, China
d
Tianjin Branch of China National Offshore Oil Company Ltd., Tianjin 300452, China
b
c

a r t i c l e

i n f o

Article history:
Received 15 September 2010
Received in revised form 15 November 2010
Accepted 27 January 2011
Available online 2 February 2011

a b s t r a c t
Three Paleogene syn-rift intervals from the Bohai Bay Basin, the most petroliferous basin in China, were
analyzed with sedimentological and geochemical techniques to characterize the lateral source rock heterogeneities, to reveal the environmental and ecological changes through geologic time and to construct
depositional models for lacustrine source rocks under different tectonic and climatic conditions. The third
(Es3) and rst (Es1) members of the Eocene Shahejie Formation and the Oligocene Dongying Formation
(Ed) display widely variable total organic carbon contents, hydrogen indices and visual kerogen compositions, suggesting changes in organic facies from deep to marginal sediments. Carefully selected deeplake facies samples from any interval, however, display fairly uniform biomarker composition. These
three intervals have distinctly different biomarker assemblages, which indicate weakly alkaline, freshwater lakes with a moderately deep thermocline during Es3 deposition, alkaline-saline lakes with shallow
chemocline during Es1 deposition and acidic, freshwater lakes with deep, unstable thermocline during
the deposition of the Dongying Formation. Such environmental changes corresponded to changes in subsidence rate and paleoclimate, from rapid subsidence and wet climate during Es3 deposition, through
slow subsidence and arid climate during Es1 deposition to rapid subsidence and wet climate during Ed
deposition and resulted in synchronous changes in terrigenous organic matter input, phytoplankton community and primary productivity. The co-evolution of environments and organisms controlled by tectonic subsidence and climate accounted for the deposition and distribution of high quality lacustrine
source rocks with distinctly different geochemical characteristics. Most rift basins experienced changes
in subsidence rates and possibly changes in climates during their syn-rift evolutions. The models constructed in this paper may have important implications for source rock prediction in other lacustrine rift
basins.
2011 Elsevier Ltd. All rights reserved.

1. Introduction
Lakes and lake deposits are important to understand and
predict because they host signicant petroleum resources (e.g.,
Kelts, 1988; Katz, 1990, 1995), are used to address climate change
and paleoclimate questions (e.g., Meyers, 1997, 2003; Fan et al.,
2007) and are signicant sources of biodiversity (Carroll and
Bohacs, 1999). Modern and ancient lakes displayed a wide variation in hydrogeology and water chemistry (e.g., Carroll and Bohacs,
Corresponding author at: State Key Laboratory of Petroleum Resources and
Prospecting, China University of Petroleum, Fuxue Road No. 18, Changping, Beijing
102249, China.
E-mail addresses: haofang@cup.edu.cn (F. Hao), zhouxh3@cnooc.com.cn
(X. Zhou), zyming@zju.edu.cn (Y. Zhu), yangyuanyuan3@foxmail.com (Y. Yang).
0146-6380/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.orggeochem.2011.01.010

1999). A number of studies of recent and ancient lake systems have


resulted in a wide variety of lacustrine source rock models for
different lakes, including the large deep anoxic lake model
(Demaison and Moore, 1980), the hypersaline lake model (Kirkland
and Evans, 1981), the oligotrophic meromictic lake model (Powell,
1986), the large mesosaline alkaline closed lake model (Kelts,
1988), the meromictic/oligomictic tropical/humid lake model
(Talbot, 1988) and the moderately deep tropical lake model (Katz,
1990).
Due to the relatively small size of the water reservoirs, lakes
have higher rates of environmental change than marine systems
(Kelts, 1988; Valero Garcs et al., 1995; Gonalves, 2002), resulting
in wide ranges of salinity and pH and marked variation of biota.
Detailed studies in recent years conrm that lacustrine source
rocks in a single rift basin may show strong vertical and horizontal

324

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

variations in hydrocarbon potentials and geochemical characteristics (e.g., Follows and Tyson, 1998; Justwan et al., 2006; Keym
et al., 2006; Hao et al., 2009a). Several case studies have been
examined to explain the horizontal (Follows and Tyson, 1998)
and vertical (Carroll and Bohacs, 1999, 2001; Bohacs et al., 2000;
Gonalves, 2002; Harris et al., 2004) heterogeneities of lacustrine
source rocks. Carroll and Bohacs (1999, 2001) and Bohacs et al.
(2000) emphasize the controls of lake types (balance lled lakes,
underlled lakes and overlled lakes) on source rock properties.

Gonalves (2002) and Harris et al. (2004), in contrast, emphasize


the role of enhanced primary productivity at late rift stages (with
relatively low subsidence rates) triggered by increased nutrient
input/recycling.
The Bohai Bay Basin is a Cenozoic lacustrine rift basin located on
the eastern coast of China (Figs. 1 and 2). The Bohai Bay Basin is the
most petroliferous basin in China, having the greatest oil production (accounting for nearly one-third of the total oil production
of China, Hao et al., 2009a), the greatest proven oil reserves and

Fig. 1. (A) Sub-basins of the Bohai Bay Basin (sub-basin classication from Allen et al. (1997)). (B) The Bozhong sub-basin showing wells from which samples were taken.
YRM = Yellow River Mouth.

325

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

BOZHONG SUB-BASIN

JIYANG SUB-BASIN

Anz

N+Q
Es

Depth (km)

Es

Ed

Ed

Pz

10
Mz

20
Mz

Post-rift Sediments
Syn-rift Sediments

Mz

10
Ek

12

15

Ek

Mz

8
Es

Ek

Pz

Ek

Ed

Pz

Es

Ek

Es

6 Mz

Pz

Es

25

Pz
0

25

Pz: Paleozoic; Mz: Mesozoic; Ek: Kongdian Formation (FM); Es:


Shahejie FM; Ed: Dongying FM; N+Q: Neogene and Quaternary

50 km

Depth (1000 ft)

Ed

N+Q

30
35

Fig. 2. Cross section showing the structural framework of the Bohai Bay Basin. Note the thick Oligocene syn-rift sediments (the Dongying Formation, Ed) and Miocene to
Quaternary post-rift sediments and strong faulting in the Bozhong sub-basin. Section location in Fig. 1A.

the greatest undiscovered oil resources in the country. The


Bozhong sub-basin, with proven in situ oil reserves greater than

Miocene

15

Neogene

10

20

Oligocene

25

30

Mingmuazhen

Pliocene
5

TECTONIC
EVOLUTION

Nmu
NmL
post-rift

Guantao

Quaternary

STRATA
LITHOLOGY
Form. Symbol
Qp
PY

Dongying

GEOLOGICAL
AGE (Ma)

35

Ng

Ed1
Ed2
Ed3

Es1

Shahejie

45

Eocene

40

Paleogene

Es2
Es3

S
syn-rift

Es4

2.5  109 tons (18.3  109 barrels) (Hao et al., 2009b; Gong et al.,
2010), is one of the most petroliferous sub-basins of the Bohai
Bay Basin. While the most important source rock interval in other
Bohai Bay sub-basins is the fourth member of the Eocene Shahejie
Formation (Es4, 50.543.0 Ma, Fig. 3) (e.g., Fuhrmann et al., 2004;
Zhang et al., 2005), Es4 has made no signicant contribution to
oil reserves so far found in the Bozhong sub-basin (Gong, 1997;
Hao et al., 2007, 2009b,c) and factors controlling the deposition
of lacustrine source rocks in the Bozhong sub-basin are not clearly
understood. All syn-rift intervals in the Bozhong sub-basin are
dominated by sandstones and mudstones/shales. The relatively
monotonous lithologic composition and paleoenvironments restrict the application of a conventional sedimentological approach
to reveal the environmental and ecological changes through geologic time (Gonalves, 2002). On the other hand, the Bozhong
sub-basin experienced multiple rifting events (Hao et al., 2009b),
and paleoclimate changed considerably during the syn-rift evolution (Wang et al., 2010). Therefore, the Bozhong sub-basin provides
an excellent natural laboratory for investigating the environmental
and ecological changes induced by changes in tectonic subsidence
and paleoclimate and for revealing the mechanisms for high quality source rock deposition under different tectonic and climatic
conditions. The purpose of our study is to assess the geochemical
variability of the Eocene and Oligocene lacustrine rift sequences
of the Bozhong sub-basin, Bohai Bay Basin, to reveal environmental
and ecological changes and to construct models for high quality
lacustrine source rocks under different tectonic and climatic conditions by integrating geological and geochemical data.
2. Geological setting

60

Paleocene

55

65

Kongdian

50

Ek

Basement
Conglomerate
Pre-Tertiary
Rocks

Sandstone

Mudstone

Potential
Source Rock

Major Oil
Reservoir

Fig. 3. Generalized stratigraphy of the Bozhong sub-basin, Bohai Bay Basin. Possible
source rock and major reservoir intervals are marked. Form. = Formation;
PY = Pingyuan.

The Bohai Bay Basin, also known as the North China Basin (Ye
et al., 1985; Hsiao et al., 2004) or the Bohai Basin (Chang, 1991;
Allen et al., 1997), is a Cenozoic lacustrine basin located on the
eastern coast of China (Fig. 1A). The Bohai Bay Basin formed on
the North China Craton (Wang and Qian, 1992; Ge and Chen,
1993) and has an area of about 200,000 km2. The mechanisms for
the formation of the Bohai Bay Basin are still controversial. Ye
et al. (1985) explained the formation of the Bohai Bay Basin with
McKenzies (1978) two stage extension model proposing Paleogene
rifting and Neogene to Quaternary thermal subsidence stages.
Allen et al. (1997, 1998), in view of the prominence of the Tan-Lu
fault, proposed a composite pull-apart basin model to explain the
formation of the Bohai Bay Basin. In recent years, more and more
workers tend to explain the Bohai Bay Basin as a Cenozoic rift basin
modied by synchronous strike slip faulting (e.g., Hsiao et al.,
2010; Qi and Yang, 2010).

326

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

The Bohai Bay Basin experienced two major tectonic evolutionary stages (Wang and Qian, 1992; Ge and Chen, 1993; Allen et al.,
1997; Hsiao et al., 2004, 2010) (Figs. 2 and 3). The basin consists of
Paleogene rifts, which were lled by a thick non-marine clastic
succession. An unconformity at the top of the syn-rift sediments
separates them from MioceneRecent strata, which were deposited during post-rift thermal subsidence (Allen et al., 1997). During
the syn-rift stage (65.024.6 Ma), a series of grabens and halfgrabens (Fig. 2) developed along major NW and NE trending fault
sets (Lu and Qi, 1997; Yang and Xu, 2004; Qi and Yang, 2010).

These grabens and half-grabens coalesced to form one large basin


during the Late Oligocene and the Bohai Bay Basin entered the
post-rift stage (24.6 Ma to the present) (Figs. 2 and 3). The Bohai
Bay Basin consists of several sub-basins, namely the Liaohe,
Bozhong, Jiyang, Huanghua, Jizhong and Linqing sub-basins (Allen
et al., 1997; Fig. 1A).
The 18,000 km2 Bozhong sub-basin, one of the six major subbasins of the Bohai Bay Basin (Allen et al., 1997), is located in the
offshore area of the Bohai Bay Basin (water depth from 5 m to
about 35 m) (Fig. 1A). The Bozhong sub-basin has the thickest

Table 1
Biomarker parameters for rock samples from different intervals in the Bozhong sub-basin, Bohai Bay Basin.
SN

Well

Depth (m)

Lithology

Interval

Pr/Ph

C35/C34

G/H

ETR

C27/C27

C19/C23

C20/C23

C24/C26

C23/C30

4MSI

C27/C29

C28/C29

S/H

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23

BZ-2
BZ-2
BZ-2
BZ-2
BZ-2
BZ-2
BZ-2
BZ-2
BZ-2
BZ-2
BZ-7
BZ-6
BZ-8
BZ-8
BZ-9
BZ-3
BZ-3
BZ-8
BZ-9
BZ-3
BZ-3
BZ-3
BZ-3

3777.5
3805
3825
3860
3905
3925
3975
4010
4122.5
4155
3205
3084.5
3171.8
3177.5
2928.5
2781
3289.5
3290
3037.5
3318
3505.5
3595.5
3691.5

Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Mudstone
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale

Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed
Ed

2.46
2.32
2.38
2.63
2.67
2.75
2.63
2.68
1.79
1.66
3.04
2.66
2.75
2.11
2.88
1.93
2.19
1.9
1.82
2.07
2.12
1.58
1.33

0.38
0.41
0.34
0.33
0.4
0.38
0.55
0.49
0.5
0.47
0.36
0.32
0.51
0.43
0.46
0.37
0.37
0.43
0.48
0.37
0.49
0.64
0.58

0.05
0.05
0.06
0.05
0.04
0.05
0.04
0.05
0.03
0.04
0.04
0.05
0.06
0.05
0.05
0.05
0.04
0.05
0.07
0.04
0.03
0.03
0.05

0.28
0.28
0.28
0.23
0.24
0.24
0.27
0.29
0.24
0.26
0.22
0.17
0.22
0.23
0.26
0.38
0.25
0.25
0.25
0.28
0.25
0.28
0.28

0.78
0.76
0.83
0.83
0.77
0.76
0.77
0.69
0.81
0.58
0.68
0.96
0.56
0.84
0.96
1.12
1.02
0.72
0.58
1.11
0.9
0.66
0.58

0.87
0.79
0.72
0.89
1.02
1.05
1.27
1.11
0.86
1.26
1.71
1.21
1.33
0.99
1.9
1.56
1.22
0.83
0.59
1.25
0.86
1.34
1.12

1.44
1.34
1.23
1.45
1.43
1.68
1.68
1.6
1.1
1.14
1.44
1.27
1.06
0.96
1.95
1.79
1.5
0.95
0.82
1.59
1.08
1.19
1.1

2.75
2.56
2.4
3.2
3.84
3.78
5.01
3.87
3.64
4.74
6.24
3.29
4.14
4.19
5.44
2.33
2.98
4.3
3.62
2.89
2.8
4.66
3.52

0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.02
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.03
0.01
0.01
0.02
0.02
0.01
0.01
0.03

0.08
0.09
0.08
0.07
0.06
0.07
0.08
0.07
0.09
0.09
0.05
0.08
0.11
0.06
0.06
0.06
0.05
0.06
0.12
0.06
0.05
0.05
0.08

0.71
0.68
0.7
0.7
0.64
0.61
0.57
0.6
0.64
0.48
0.85
0.94
0.59
0.92
0.74
0.84
1
0.88
0.93
0.96
0.88
0.59
0.75

0.4
0.4
0.43
0.43
0.44
0.43
0.39
0.44
0.58
0.52
0.47
0.51
0.57
0.69
0.3
0.43
0.6
0.66
0.84
0.59
0.58
0.52
0.64

0.08
0.07
0.07
0.07
0.06
0.06
0.07
0.08
0.08
0.08
0.13
0.18
0.09
0.19
0.14
0.18
0.11
0.15
0.24
0.11
0.08
0.08
0.14

24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40

BZ-5
BZ-5
BZ-5
BZ-5
BZ-5
BZ-5
BZ-5
BZ-7
BZ-6
BZ-8
BZ-9
BZ-11
BZ-11
BZ-8
BZ-9
BZ-10
BZ-4

4401
4407
4423.5
4449
4459.5
4476
4492.5
3287.5
3264
3417.5
3185
3033
3039
3450
3240
3250
3646

Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale

Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1
Es1

1.21
1.26
1.3
1.69
1.38
1.22
1.35
1.27
1.48
1.3
1.92
1.39
1.81
1.16
1.49
1.33
1.47

0.79
0.62
0.68
0.54
0.57
0.58
1.02
0.47
0.54
0.46
0.48
0.67
0.56
0.49
0.5
0.57
0.6

0.4
0.46
0.41
0.26
0.44
0.41
0.42
0.53
0.07
0.14
0.08
0.2
0.24
0.31
0.22
0.64
0.16

0.67
0.65
0.62
0.43
0.59
0.62
0.64
0.45
0.36
0.35
0.25
0.62
0.43
0.42
0.39
0.52
0.51

0.28
0.22
0.29
0.51
0.25
0.25
0.24
0.29
0.61
0.49
0.74
0.53
0.99
0.46
0.61
0.6
0.47

0.24
0.34
0.27
0.65
0.43
0.22
0.31
0.17
0.3
0.34
0.53
0.15
0.53
0.24
0.26
0.18
0.47

0.69
0.8
0.78
1.04
1.06
0.75
1.08
0.5
0.49
0.57
0.72
0.43
1.06
0.5
0.54
0.45
1.09

0.59
0.59
0.55
0.73
0.61
0.59
0.55
0.92
1.23
1.75
2.31
0.53
1.23
1.06
1.05
0.77
0.8

0.29
0.14
0.15
0.09
0.12
0.18
0.18
0.05
0.03
0.02
0.02
0.14
0.08
0.04
0.04
0.06
0.22

0.19
0.23
0.22
0.17
0.23
0.23
0.22
0.34
0.18
0.11
0.1
0.23
0.25
0.16
0.13
0.21
0.18

0.98
0.86
1.02
0.89
0.83
0.9
0.99
0.86
0.85
0.91
0.93
1.1
1.41
0.98
1.11
1.09
0.97

0.88
0.91
0.88
0.85
0.84
0.84
0.85
0.74
0.57
0.7
0.8
0.57
0.42
0.79
0.92
0.92
0.7

0.44
0.39
0.39
0.31
0.33
0.41
0.42
0.18
0.16
0.18
0.2
0.57
0.28
0.22
0.24
0.31
0.2

41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57

BZ-8
BZ-8
BZ-8
BZ-9
BZ-9
BZ-9
BZ-4
BZ-4
BZ-4
BZ-1
BZ-1
BZ-1
BZ-7
BZ-7
BZ-7
BZ-1
BZ-1

3555
3570
3585
3537.5
3587.5
3637.5
3745
3764.5
3775
3632.5
3662.5
3687.5
3335
3598.5
3997.3
3945
3985

Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale
Shale

Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3
Es3

1.64
1.57
1.47
1.44
1.62
1.37
1.55
1.59
1.48
2.05
1.87
1.7
1.51
1.67
1.71
1.81
1.55

0.54
0.55
0.6
0.54
0.56
0.53
0.53
0.52
0.53
0.46
0.43
0.4
0.54
0.62
0.42
0.48
0.43

0.07
0.07
0.08
0.07
0.07
0.1
0.04
0.05
0.06
0.06
0.06
0.06
0.05
0.07
0.14
0.08
0.09

0.26
0.26
0.31
0.28
0.29
0.3
0.5
0.43
0.39
0.16
0.29
0.31
0.33
0.32
0.54
0.37
0.4

0.67
0.69
0.64
0.72
0.54
0.55
0.2
0.26
0.29
0.95
0.95
0.94
0.53
0.66
0.81
0.84
0.83

0.54
0.54
0.42
0.64
0.67
0.39
0.37
0.36
0.37
0.65
0.56
0.38
0.42
1.01
0.24
0.5
0.48

0.69
0.66
0.56
0.74
0.79
0.6
0.82
0.83
0.79
0.83
0.75
0.64
0.64
0.62
0.44
0.77
0.9

1.59
1.24
1.05
1.55
1.35
0.7
1.14
1.27
1.08
2.74
2.46
1.91
1.24
1.53
0.09
1.71
1.46

0.03
0.03
0.04
0.03
0.05
0.08
0.03
0.03
0.03
0.02
0.02
0.03
0.03
0.02

0.15
0.2
0.39
0.09
0.16
0.52
0.16
0.56
0.45
0.09
0.12
0.13
0.34
0.29
0.15
0.13
0.13

0.81
0.74
0.71
0.81
0.8
0.61
0.73
0.71
0.67
0.9
0.99
0.96
0.97
0.42
0.52
0.83
0.75

0.64
0.63
0.7
0.69
0.7
0.82
0.51
0.57
0.56
0.52
0.74
0.71
0.62
0.5
0.57
0.58
0.49

0.11
0.11
0.11
0.11
0.13
0.16
0.1
0.1
0.11
0.19
0.16
0.16
0.16
0.13
0.37
0.22
0.23

0.08
0.07

Note: SN = sample number; Pr/Ph = pristane/phytane; C35/C34 = C35 22S/C34 22S hopane; G/H = gammacerane/ C30 hopane; ETR = (C28 + C29)/(C28 + C29 + Ts); C27/C27 = C27
(20R + 20S) diasteranes/C27 (20R + 20S) steranes; C19/C23 = C19 tricyclic terpane/C23 tricyclic terpane; C20/C23 = C20 tricyclic terpane/C23 tricyclic terpane; C24/C26 = C24 tetracyclic terpane/C26 tricyclic terpane; C23/C30 = C23 tricyclic terpane/ C30 hopane; 4MSI = 4-methylsterane index (4-methylsteranes/C29 steranes); C27/C29 = C27/C29 steranes;
C28/C29 = C28/C29 steranes; S/H = steranes/hopanes (C27 C29 steranes/C27 C35 hopanes).

327

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

Gas chromatographic (GC) analyses of the saturated hydrocarbon


fractions were achieved using a HP6890 chromatograph equipped
with a PONA fused silica column (60 m  0.20 mm i.d., lm thickness 0.5 lm). The oven temperature was initially held at 35 C for
5 min, programmed to 80 C at 2 C/min, to 300 C at 4 C/min and
held at 300 C for 30 min. Helium was used as the carrier gas. GC
mass spectrometry (GCMS) analyses of the saturate fractions
were performed with a HP6890GC/5973MSD instrument equipped
with a HP-5MS fused silica column (30 m  0.25 mm i.d., lm
thickness 0.25 lm). The GC oven temperature for analysis of the
saturate fractions was initially held at 50 C for 2 min, programmed to 100 C at 20 C/min and to 310 C at 3 C/min, and
held at 310 C for 16.5 min. Biomarker ratios were calculated from
peak areas of individual compounds. Kerogen was isolated from
extracted samples by overnight treatment with 6 N HCl followed
by concentrated HF.

Tertiary and Quaternary sediments (up to 11 km, Fig. 2) in the


Bohai Bay Basin. The thick Oligocene syn-rift sediments (the
Dongying Formation) and Miocene to Quaternary post-rift sediments, as well as strong late stage faulting made the Bozhong
sub-basin distinctly different from other Bohai Bay sub-basins
(Fig. 2). The syn-rift sediments are composed of the Paleocene
Eocene Kongdian Formation (Ek), the Eocene Shahejie Formation
(Es) and the Oligocene Dongying Formation (Ed) (Figs. 2 and 3).
These formations were restricted to the grabens and half-grabens
(Fig. 2) and were deposited in uvial-lacustrine environments
(Allen et al., 1997, 1998; Gong, 1997). The post-rift sediments consist of the Miocene Guantao Formation (Ng), MiocenePliocene
Minghuazhen Formation (Nm) and the Quaternary Pingyuan Formation (Qp) (Fig. 3). These formations are widespread (Fig. 2), and
are dominated by uvial deposits (Gong, 1997; Yang and Xu, 2004).
The Bozhong sub-basin experienced multiple rifting events,
which caused signicant variations in syn-rift subsidencesedimentation rates over time. The subsidence rate was up to
600 m/Ma during the deposition of the third member of the Eocene
Shahejie Formation (Es3), decreased to <300 m/Ma during the
deposition of the rst member of the Eocene Shahejie Formation
(Es1), and increased to up to 600 m/Ma again during the deposition
of the Oligocene Dongying Formation (Ed). On the other hand,
paleoclimate changed considerably during the syn-rift evolution
of the Bohai Bay Basin (Wang et al., 2010). Wet, northern subtropical climate dominated during the deposition of Es3. Aridity
increased from the late stage of Es3 deposition to Es2 and Es1
deposition stage and decreased thereafter, with wet, northern subtropical climate dominating during the deposition of the Oligocene
Dongying Formation (Wang et al., 2010).

4. Results
4.1. RockEval pyrolysis
RockEval pyrolysis is a commonly used technique to classify
organic matter (OM) types and to assess hydrocarbon generating
potentials (e.g., Peters, 1986). The third member of the Eocene
Shahejie Formation (Es3, 43.038.0 Ma) has total organic carbon
(TOC) contents of 0.19.19% and RockEval S2 values of 0.02
63.08 mg HC/g rock. Hydrogen indices for Es3 samples range from
<50 to 1115 mg HC/g TOC, suggesting different organic matter
types (Fig. 4A). The rst member of the Shahejie Formation (Es1,
35.832.8 Ma) displays TOC contents of 0.226.80%, RockEval S2
values of 0.2550 mg HC/g rock, and hydrogen indices of 15
777 mg HC/g TOC (Fig. 4B). The Oligocene Dongying Formation
(Ed, 32.824.6 Ma) has TOC contents ranging from 0.35% to
3.91%, RockEval S2 values from 0.126.3 mg HC/g rock and hydrogen indices from <50 to 716 mg HC/g TOC (Fig. 4C). Most samples
have RockEval Tmax lower than 445 C, suggesting thermal maturity from immature to early oil generation (Peters, 1986).

3. Samples and methods


Two sample sets were analyzed in this study. One sample set
consisted of more than 300 mudstone/shale samples of different
sedimentary facies. This sample set was used for bulk geochemistry analysis such as RockEval pyrolysis, visual kerogen observation and gas chromatography of saturated hydrocarbons to reveal
changes in organic facies within a stratigraphic interval, from deep
to lake margin facies. Another sample set consisted of 57 deep lake
shales/mudstones (Table 1). This sample set was used for gas
chromatographymass spectrometry (GCMS) analysis to investigate environmental and ecological changes in the lakes through
geologic time.
All rock samples were cleaned prior to powdering. Soxhlet
extraction was conducted using chloroform/methanol (87:13) for
72 h and the isolated extractable organic matter was separated
into saturated hydrocarbons, aromatic hydrocarbons and polars.

In accordance with the wide variations in TOC contents, Rock


Eval S2 peaks and hydrogen indices (Fig. 4), samples from different
intervals show wide variations in visual kerogen compositions
(Fig. 5A and B). Kerogens from Es3 display woody organic matter
contents ranging from <133%, and amorphous organic matter
(AOM) contents from 577%. Kerogens from Es1 have woody
organic matter contents of 039%, and AOM contents of 0.578%
(Fig. 5). Kerogens from the Dongying Formation have woody

1000

1000

1000
Hydrogen Index (mg HC/g TOC)

4.2. Visual kerogen analysis

(B) Es1

(A) Es3
I

800

Ro=0.5%

(C) Ed

800

Ro=0.5%

II

600

600

600

400

400

400

200

Ro=1.0%

III
420

Ro=0.5%

II

II

0
400

800

440

460

Tmax (C)

480

500

200

Ro=1.0%

III
0
400

420

440

460

Tmax (C)

480

500

200

Ro=1.0%

III
0
400

420

440

460

480

500

Tmax (C)

Fig. 4. Variation of hydrogen index as a function of Tmax for samples from different possible source rock intervals in the Bozhong sub-basin, Bohai Bay Basin.

328

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

1000

(A)
800

Hydrogen Index (mg HC/g TOC)

Hydrogen Index (mg HC/g TOC)

1000

Ed
Ed Rock
Rock
Es
Rock
Es11Rock
Es
Es33Rock
Rock

600
400
200
0

10

20
30
40
Woody Organic Matter (%)

800
600
400
200
0

50

10

20
30
40
50
60
Amorphous Organic Matter (%)

70

80

800
Hydrogen Index (mg HC/g TOC)

800
Hydrogen Index (mg HC/g TOC)

(B)

(C)

700
600
500
400
300
200
100
0
0.0

0.5

1.0
1.5
2.0
2.5
3.0
Pristane/Phytane (Pr/Ph)

3.5

(D)

700
600
500

Phytane from
methanogenic archaea?

400
300
200
100
0

4.0

3
4
Phytane/nC18

Fig. 5. Variation of hydrogen indices as a function of woody organic matter contents (A), amorphous organic matter contents (B), Pristane/Phytane ratios (C) and Phytane/nC18 ratios (D), showing source rock heterogeneities in the Bozhong sub-basin, Bohai Bay Basin.

(B)
Ed
Ed Rock
Rock
Es
Es11Rock
Rock
Rock
Es
3
Es3 Rock

Phytane/n-C18

Pristane/n-C17

(A)

0
400

410

420

430
Tmax (C)

440

450

0
400

460

410

420

430
Tmax (C)

440

10

10

(C)

(D)
Ox

Pristane/n-C17

s t
ce e n
ur nm
t
so iro
en
v
OM n OM nm
d le
o
xe ona lgal vir
i
M it i
A en
g
ns
c in
t ra
du
e
r

g
in

10

1.0

460

c
du
Re

1.0
Phytane/n-C18

0.1
0.1

Te

n
izi

Pristane/Phytane

id

1.0

sO
ou
en
g
i
rr

450

0.1
0.1

1.0
Phytane/n-C18

10

Fig. 6. Variation of pristane/phytane (A) and Pristane/n-C17 (B) as a function of Phytane/n-C18, and variation of Pristane/n-C17 (C) and Phytane/n-C18 (D) as a function of Tmax
for rock samples from the Bozhong sub-basin, Bohai Bay Basin.

329

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

(A) Well BZ-3, 3505.5m, Ed (Sample 21)


m/z 191

m/z 191

m/z 217

C24 Tet
C30 hopane

C20
C19

C27

C29

C27 RA
steranes

C21

C23

C28
4-MS

C24
C22

Gam
C25 C26

(B) Well BZ-7, 3287.5m, Es 1 (Sample 31)


C27

C23

C21

C28
C24

C20
C19

C25 C26

C22

C29

C24 Tet
4-MS

Gam

(C) Well BZ-4, 3764.5m, Es 3 (Sample 48)


C29
C21

C23

C20
C19

C27
C24 Tet

C22

C25 C26

4-MS
C28

C24
Gam

Fig. 7. Representative mass chromatograms of terpane (m/z = 191) and sterane (m/z = 217) series of saturate fractions for different stratigraphic intervals in the Bozhong subbasin, Bohai Bay Basin. Peaks marked by solid dots are tricyclic terpanes (C19C26). C24 Tet = C24 tetracyclic terpane; Gam = gammacerane; RA = re-arranged; C27, C28 and C29
represent C27 sterane 20R, C28 sterane 20R and C29 sterane 20R, respectively; 4-MS = 4-methylsteranes. Note the abundant C19, C20 tricyclic terpanes, C24 tetracyclic terpane
and C27 re-arranged steranes for Ed sample (A), abundant gammacerane for Es1 sample (B), and abundant 4-methylsteranes for the Es3 sample (C).

organic matter contents ranging from <139%, and AOM contents


from <0.1% to 76% (Fig. 5).
4.3. GC and GCMS analysis
GC analyses were carried out on more than 300 samples. Es3
displays pristane/phytane (Pr/Ph) between 0.2 and 3.21, Pr/n-C17
between 0.31 and 4.95 and Ph/n-C18 between 0.21 and 5.53 (Figs. 5
and 6). Es1 has Pr/Ph ranging from 0.163.55, Pr/n-C17 from 0.18
1.76, and Ph/n-C18 from 0.174.59 (Figs. 5 and 6). The Dongying
Formation displays Pr/Ph of 0.593.29, Pr/n-C17 of 0.282.76, and
Ph/n-C18 of 0.132.64 (Figs. 5 and 6).
GCMS analyses were conducted on 57 laminated deep lake
samples and representative m/z 191 and m/z 217 chromatograms
for the analyzed samples are displayed in Fig. 7. The analyzed samples display wide variations in the relative abundances of C19, C20
and C23 tricyclic terpanes, C24 tetracyclic terpane, gammacerane,
and 4-methylsteranes (Fig. 7, Table 1), suggesting obvious changes
in organic matter origins and/or depositional environments.
5. Discussion
5.1. Source rock heterogeneity of different stratigraphic intervals
The third (Es3) and rst (Es1) members of the Eocene Shahejie
Formation and the Oligocene Dongying Formation (Ed) are potential source rocks in the Bozhong sub-basin (Hao et al., 2009a,b;
Gong et al., 2010). It is therefore of great signicance to decipher
the heterogeneity of each interval both for petroleum reserve
assessment and for oilsource rock correlation (e.g., Justwan
et al., 2006; Keym et al., 2006; Curiale, 2008). Es3, Es1 and Ed all
display wide variations in RockEval hydrogen indices (Fig. 4). It
should be pointed out that most samples have RockEval Tmax

lower than 445 C (Fig. 4). Therefore, the wide variations in


hydrogen indices could not have been caused by maturity changes.
Predictably, the RockEval hydrogen indices decrease as woody
organic matter contents increase (Fig. 5A). Several case studies
showed that amorphous organic matter (AOM) was derived mainly
from algal organic matter deposited into reducing environments
and therefore was rich in hydrogen (e.g., Powell et al., 1990; Tyson,
1995; Ercegovac and Kostic, 2006). In the Bozhong sub-basin, however, there is a loose correlation between RockEval hydrogen indices and AOM contents and samples with similar AOM contents
display wide variation in hydrogen indices (Fig. 5B). Noticeably,
the Dongying Formation samples usually have lower hydrogen
indices than Es3 and Es1 samples with similar AOM contents
(Fig. 5B), suggesting that AOM in the Dongying Formation experienced more intensive degradation during or immediately after
deposition (Frimmel et al., 2004). Several Es1 and Es3 samples dominated by AOM have hydrogen indices no higher than 250 mg HC/g
TOC (Fig. 5B), suggesting that AOM in these samples is poor in
hydrogen. The loose correlation between hydrogen indices and
AOM contents and the low hydrogen indices for samples dominated by AOM seem to indicate that AOM in the Bozhong sub-basin
is complex in hydrogen contents. This is consistent with the observation of Ebukanson and Kinghorn (1985) and Hao et al. (1993)
who concluded that AOM could be formed both from algal organic
matter and from higher plant organic matter and therefore could
be either hydrogen rich or hydrogen poor.
RockEval hydrogen indices decrease with increasing pristane/
phytane (Pr/Ph) ratios (Fig. 5C), and high abundance of phytane (Ph/n-C18 > 2.0) always occurs in samples with hydrogen
indices >500 mg HC/g TOC (Fig. 5D). Because most samples have
RockEval Tmax < 445 C (Fig. 4) and Pr/n-C17 and Ph/n-C18 show
no obvious correlation with Tmax (Fig. 6A and B), the overall decrease of hydrogen indices with Pr/Ph is caused mainly by changes
in redox conditions during or immediately after deposition of

330

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

0.3 0.6
0.5 1.0 0

C28/C29
C27/C29

0.8 1.6 0
0.4 0.8 0

4MSI
C27/C27

4
1
1
0.6 1.2 0
0.4 0.8 0

ETR
G/H
C35/C34

0.4 0.8 0
2

4600

4400

4200

4000

3800

3600

3400

3200

3000

2800

2600

Pr/Ph

0.6 1.2 0

C19/C23

C20/C23

C24/C26

C23/C30

(Seplveda et al., 2009), or sulfurization of functional lipids during


early diagenesis and subsequent temperature cleavage of the sulfur bonds during catagenesis in sulfur rich rocks (e.g., Keely
et al., 1993; Koopmans et al., 1996). Cenozoic rocks in the Bozhong
sub-basin are sulfur poor and the sulfurizationcleavage mechanism

0.2 0.4 0

S/H

source sediments, rather than by changes in maturity levels. Three


mechanisms have been proposed in the literature to explain the
abnormally high abundance of phytane (high Ph/n-C18 ratios):
contribution from methanogenic archaea (e.g. Rowland, 1990;
Fuhrmann et al., 2004), contribution from eukaryotic phytoplankton

Depth (m)
Fig. 8. Variations of major biomarker parameters reecting depositional conditions and/or organic matter input as a function of burial depth. Open triangles: Ed samples;
Crosses: Es1 samples; solid dots: Es3 samples. Abbreviations for biomarker parameters are explained in Table 1.

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

is therefore not applicable. Contribution from eukaryotic phytoplankton may simultaneously lead to high abundances of both
pristane and phytane, as is the case in the Levant Platform of
central Jordan (Seplveda et al., 2009). Yet in the Bozhong subbasin Pr/Ph ratios decrease as Ph/n-C18 increases (Fig. 6C) and
samples with Ph/n-C18 greater than 2.0 display Pr/n-C17 ratios
obviously lower than the Pr/n-C17 ratios for many Dongying
Formation samples that have Ph/n-C18 < 1.0 (Fig. 6D). Therefore,
contribution from eukaryotic phytoplankton could not be the main
cause of the high phytane abundance observed in the Bozhong subbasin. In the Bozhong sub-basin, samples with high Ph/n-C18 ratios
have a large isotopic difference between saturate and aromatic
fractions and plot above the best t line (d13CAro = 1.14d13CSat + 5.46) on the so called Sofer-plot (Sofer, 1984) (Guo, 2009,
personal communication). The 13C depleted saturate fractions resulted from input from methanotrophic bacteria (Collister and
Wavrek, 1996). Methanotrophic bacteria thrive at the chemocline
of stratied lakes (Collister and Wavrek, 1996). The 13C depleted
saturate fractions as a result of methanotrophic input of isotopically depleted lipids imply that a considerable amount of methane
should have been generated below the chemocline. Therefore, we
explain the high Ph/n-C18 ratios for samples with high hydrogen
indices as reecting a contribution from methanogenic archaea.
Based on RockEval hydrogen indices (Fig. 4), elemental compositions of kerogens (not shown), visual kerogen compositions
(Fig. 5A and B) and redox conditions reected by isoprenoid hydrocarbon distributions (Figs. 5C, D and 6), four types of organic facies
(Jones, 1987) can be recognized. Organic facies A has hydrogen
indices >700 mg HC/g TOC and is dominated by uorescent AOM
and well preserved algal materials usually with low Pr/Ph (<1.0)
and high Ph/n-C18 (>1.0 to 1.5) ratios. Organic facies B usually displays hydrogen indices between 500 and 700 mg HC/g TOC. Organic
facies B is also dominated by AOM but may have woody organic
matter contents up to 15%. Organic facies BC has hydrogen indices
between 200 and 500 mg HC/g TOC and woody organic matter
contents of 1045%. Organic facies C has hydrogen indices
<200 mg HC/g TOC and is dominated by woody organic matter.
AOM in organic facies C is ne grained and is red in color in transmitted light and shows no uorescence in ultraviolet light, which is
distinctly different from AOM in organic facies A and B.
Both Es3 and Es1 contain organic facies AC. The Dongying Formation contains organic facies B to C without well developed organic facies A. The wide, successive variations in hydrogen
indices (Fig. 4) and visual kerogen composition (Fig. 5A and B) suggest strong heterogeneities of all the three intervals, from hydrogen rich organic facies in the deep lake sediments to hydrogen
poor facies in lake marginal sediments. Such heterogeneities may
be explained as reecting varying redox conditions from anoxic
through transitional to oxic (Fig. 6D). It should be pointed out that,
in the Bozhong sub-basin, organic facies BC and C usually have total organic carbon (TOC) contents <1.0% (Hao et al., 2010). The relatively low hydrogen indices and TOC contents mean that the
hydrocarbon generating potential of organic facies BC and C is
much lower than that of organic facies A and B. In addition, organic
facies BC and C might contribute to condensates and hydrocarbon
gases, but are of minor signicance for normal oils due to the
oxygen rich nature of the organic matter (Price, 1989). Therefore,
to understand the mechanisms for the deposition of organic facies
A and B and to predict their distribution are of great signicance
for petroleum exploration in lacustrine basins.
5.2. Changes in lake water chemistry
Primary productivity and redox conditions are among the most
important factors controlling high quality source rock (organic facies A and B) deposition (e.g., Kelts, 1988; Huc et al., 1992; Katz,

331

1990, 2005; Tyson, 2005). The primary productivity and redox conditions in a lake are largely controlled by lake water chemistry,
such as salinity and acidity/alkalinity. Rapid progress in organic
geochemistry in the last 20 years makes it possible to qualitatively
estimate changes in lake water chemistry from biomarker parameters (e.g., Volkman et al., 1998; Peters et al., 2005, p. 483580).
The great thickness of the syn-rift succession in the Bozhong
sub-basin (cf. Fig. 2) makes it impossible to construct a biomarker
prole in a single well. In order to reveal changes in lake water
chemistry and the resultant changes in biota through geologic
time, 57 samples of deep-lake facies from different intervals were
analyzed (Table 1) and a composite biomarker prole (Figs. 8 and
9) was constructed. Hao et al. (2010) showed that no terpane
and sterane parameters for these samples displayed any obvious
correlation with the C29 bb/(bb + aa) sterane ratio (an effective
maturity parameter believed to be independent of organic matter
input; Peters et al., 2005, p. 625630). In addition, although the
selected sample set covers a relatively wide depth range (2781
4492.5 m), most parameters show no obvious correlation with
burial depth (Fig. 8), and the variation trends of several parameters
are even opposite to those caused by thermal maturity. For
instance, if Pr/Ph and C19/C23 TT had been intensively inuenced
by thermal maturity, they would have had to increase with
increasing depth. Yet, exactly the opposite is observed (Fig. 8). It
appears that all parameters displayed in Figs. 8 and 9 are not intensively inuenced by thermal maturity and therefore can be used to
reect organic matter input and/or depositional conditions. Carefully selecting deep-lake facies samples based on the result of
sedimentological analysis from different wells and constructing a
composite biomarker prole could minimize the inuence of local
water inow and/or sediment input. Despite the fact that samples
for any interval were from different wells and cover a relatively
wide depth range (Table 1), all three intervals display relatively
narrow variation ranges for most parameters (Fig. 9). This on one
hand suggests that the inuence of local water inow and/or sediment input, if any, is minor, and on the other hand implies that
these parameters have not been signicantly affected by thermal
maturity.
Pr/Ph and C35 22S/C34 22S hopane are effective oxicity parameters (see Peters et al., 2005, p. 499502, 566569 and references
therein). Es1 has the lowest Pr/Ph ratios and the highest C35 22S/
C34 22S hopane ratios (Table 1, Fig. 9A and B). In contrast, the
Dongying Formation displays the highest Pr/Ph ratios and the
lowest C35 22S/C34 22S hopane ratios. The relatively low Pr/Ph
ratios and high C35 22S/C34 22S hopane ratios for Es3 and Es1 samples suggest that anoxic conditions prevailed in the bottom water
during Es3 and Es1 deposition, whereas less reducing conditions
dominated in the bottom water in lakes during Ed deposition. This
is consistent with the result of bulk geochemistry analysis. On the
Pr/n-C17 vs Ph/n-C18 plot, many Es3 and Es1 samples plot in areas
suggesting euxinic, anoxic conditions, whereas most Ed samples
fall in the transitional to oxic elds (Fig. 6D).
The gammacerane index (gammacerane/ab C30 hopane), extended tricyclic terpane ratio [ETR = (C28 + C29)/(C28 + C29 + Ts)]
and C27 ba (20R + 20S) diasteranes/C27 abb (20R + 20S) sterane
(C27 Dia/C27 ST) were used to reect changes in water salinity/alkalinity through geologic time (Fig. 9CE).
Gammacerane is believed to form by reduction of tetrahymanol
(e.g., Venkatesan, 1989; ten Haven et al., 1989). The principal
source of tetrahymanol appears to be bacterivorous ciliates, which
occur at the interface between oxic and anoxic zones in stratied
water columns (Sinninghe Damst et al., 1995). Therefore, abundant gammacerane is usually believed to indicate the presence of
a stratied water column (e.g., Sinninghe Damst et al., 1995;
Peters et al., 2005, p. 576; Seplveda et al., 2009). Although a
stratied water column can result from either hypersalinity or a

(L)

0.5

(K)

0.8 1.6 0

(J)

0.3 0.6 0

(I)

0.2 0.4 0

(H)

(G)

(F)

(E)

0.6 1.2 0

(D)

0.4 0.8 0

(C)

0.4 0.8 0

(B)

0.6 1.2 0

(A)

55

50

40

35

30

20

15

10

25

Es2+1 (38.032.8 Ma)

45

Es3 (43.038.0 Ma)

SN

Pr/Ph

4 0

C35/C34

G/H

ETR

C27/C27

C19/C23

2 0

C20/C23

2 0

C24/C26

C23/C30

4MSI

C27/C29

C28/C29

1 0

(M)

0.3 0.6

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

S/H

332

E3 (32.824.6 Ma)

Fig. 9. Composite organic geochemistry prole showing changes in major biomarker parameters reecting depositional conditions and/or organic matter input from Es3
through Es1 to Ed3. Vertical lines represent the average values for the interval. SN = sample number. Abbreviations for biomarker parameters are explained in Table 1.

temperature gradient (e.g., Bohacs et al., 2000), high abundances of


gammacerane are mostly found in evaporite or high salinity
environments (Fu et al., 1990; Chen et al., 1996; Ritts et al.,
1999; Hanson et al., 2000, 2001; Holba et al., 2003; Summons
et al., 2008).

Holba et al. (2001) used ETR to differentiate crude oils generated from Triassic, Lower Jurassic and MiddleUpper Jurassic
marine source rocks. They observed a sharp drop in ETR at the
end of the Triassic that corresponds to a major mass extinction
and implied that the mass extinction may have had an impact on

333

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

the principal biological sources of tricyclic terpanes. The effectiveness of ETR as an age related parameter was questioned by Ohm
et al. (2008). In the Bozhong sub-basin, ETR increases with increasing gammacerane/ab C30 hopane and sterane/hopane ratios and
decreases with increasing Pr/Ph ratios (Fig. 10AC). Similar trends
were also observed in lacustrine oils and source rocks in the
Junggar Basin where high ETR occurs in crude oils and source rocks
with abundant b-carotane and ETR increases as gammacerane/ab
C30 hopane increases (Hao et al., in press). Our observations in
the Junggar and Bohai Bay basins are consistent with those of
Kruge et al. (1990a,b) and De Grande et al. (1993) who concluded
that fossil lipids of prokaryotes in saline, alkaline lakes are rich in
precursors of extended tricyclic terpanes. The close correlation of
ETR with Pr/Ph, sterane/hopane and gammacerane/ab C30 hopane
ratios indicate that in lacustrine environments, ETR is an effective

0.8

0.8

(A)
0.6

0.4
Ed
Ed Rock
Rock
Es
Es11Rock
Rock
Es
Es33Rock
Rock

0.2
0.0
0.03

0.1

0.4
0.2
0.0
0.0

1 .0

(C)

0.6

ETR

ETR

ETR

0.8

(B)

0.6

0.4
0.2

0.1

0.2

0.3

0.4

0.5

0.0
1.0

0.6

1.5

2.0

S/H

G/H
1.5

(D)
C27/C27

1.0

0.5

2.5

3.0

3.5

Pr/Ph
7.0

(E)

(F)

6.0
5.0

1.0

C24/C26

1.5

C27/C27

indicator of the salinity and alkalinity during or immediately after


deposition of source sediments (Hao et al., 2009a).
Bennett and Olsen (2007) used C27 Dia/C27 ST as an indicator of
source rock lithology, and showed that carbonate source rocks had
low C27 Dia/C27 ST ratios. All analyzed samples are lacustrine shale/
mudstone (Table 1) and therefore variations in the ratio cannot reect lithologic changes. Yet the analyzed samples display a wide
range of C27 Dia/C27 ST ratios (0.201.12, Table 1, Fig. 9E). C27
Dia/C27 ST increases as Pr/Ph increases for samples with Pr/
Ph < 2.0, but remains constant at relatively high values for samples
with Pr/Ph > 2.0 (Fig. 10D). With the exception of three samples,
C27 Dia/C27 ST decreases as the gammacerane index increases
(Fig. 10E). Acidic and oxic conditions facilitate diasterene formation during diagenesis (Brincat and Abbott, 2001; Peters et al.,
2005, p. 533), therefore, variations in C27 Dia/C27 ST ratios in

0.5

4.0
3.0
2.0
1.0

0.0
1.0

1.5

2.0

2.5

3.0

0.0
0.0

3.5

0.2

0.4

Pr/Ph
60

3.0
2.0
1.0
2.5

3.0

50
40
30
20
0.0

3.5

C29 Steranes (%)

4.0

2.0

1.0

0.5

1.0

Pr/Ph

1.5

50
40
30
20
0.0

2.0

2.0

4.0

6.0

8.0

C24/C26

0.6

(J)

2.0

(I)

C19/C23

1.0

1.5

60

(H)
C29 Steranes (%)

5.0

1.5

0.5

C19/C23

(G)

6.0

C24/C26

0.0
0.0

0.8

G/H

7.0

0.0
1.0

0.6

0.6

(K)

(L)

0.4

S/H

0.4

S/H

C28/C29

0.8
0.6

0.2

0.2

0.4
0.2
0.4

0.6

0.8

1.0

C27/C29

1.2

1.4

1.6

0.0
0.1

0.2

0.3

C28 Steranes (%)

0.4

0.0
0.3

0.4

0.5

0.6

C29 Steranes (%)

Fig. 10. Correlation between various biomarker parameters reecting organic matter input and/or depositional environments in the Bozhong sub-basin, showing the
differences in biomarker compositions among the three source rock intervals. Abbreviations for biomarker parameters are explained in Table 1.

334

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

non-marine clastic successions reect changes in Eh and pH of the


lake systems.
Es1 has the highest gammacerane index, highest ETR but the
lowest C27 Dia/C27 ST (Table 1, Figs. 7, 9CE), suggesting saline,
alkaline lakes. This is supported by the fact that relatively abundant carbonates including dolomites were deposited in the shallow
lake facies of Es1. Dolomites are known to form in concentrated
alkaline lakes, but are absent from freshwater lakes (Wolfbauer
and Surdam, 1974; Jones and Bowser, 1978; Fuhrmann et al.,
2004). In contrast, the Dongying Formation displays the lowest
gammacerane index, lowest ETR but the highest C27 Dia/C27 ST
(Table 1, Figs. 7, 9CE), suggesting freshwater, acidic lakes. Es3
has medium gammacerane index, ETR and C27 Dia/C27 ST (Table 1,
Fig. 9CE). We interpreted lakes during Es3 deposition as fresh to
brackish water, weak alkaline lakes.
In summary, considerable changes in lake water chemistry occurred from the middle Eocene to early Oligocene (from Es3 deposition to Ed deposition) in the Bozhong sub-basin. Fresh to
brackish, weakly alkaline lakes with anoxic bottom water conditions dominated during Es3 deposition. Saline, alkaline lakes with
euxinic bottom water conditions occurred during Es1 deposition.
In contrast, freshwater, acidic lakes with dysoxic bottom water
conditions prevailed during the deposition of the Dongying
Formation.
5.3. Changes in terrigenous organic matter input and primary
producers
Terpane parameters have the potential to reect terrigenous organic matter input (e.g., Peters et al., 2005, p. 538580). High C19
tricyclic terpane/C23 tricyclic terpane (C19/C23 TT) and C20 tricyclic
terpane/C23 tricyclic terpane (C20/C23 TT) ratios indicate important
contribution from terrigenous organic matter (e.g., Hanson et al.,
2000; Preston and Edwards, 2000; George et al., 2004; Volk et al.,
2005; Hao et al., 2009a). In the Bohai Bay Basin, the C24 tetracyclic
terpane/C26 tricyclic terpane (C24 Tet/C26 TT) ratios increase as C19/
C23 TT and Pr/Ph ratios increase (Fig. 10F,G), indicating that high
abundances of C24 tetracyclic terpane (high C24 Tet/C26 TT ratios)
are diagnostic of terrigenous organic matter input in lacustrine
systems (Philp and Gilbert, 1986; Hanson et al., 2000; Bohacs
et al., 2000; George et al., 2004). Es1 has the lowest C19/C23 TT,
C20/C23 TT and C24 Tet/C26 TT ratios but the highest C23 tricyclic terpane/ab C30 hopane (C23 TT/C30 H) ratios (Table 1, Figs. 7, 9FI),
indicating no or minor contributions from terrigenous organic
matter. This is consistent with the strongly reducing conditions
and well developed water column stratication reected by the
low Pr/Ph ratios and high C35 22S/C34 22S hopane, gammacerane/
ab C30 hopane and ETR (Fig. 9AD). In contrast, the Dongying Formation displays the highest C19/C23 TT, C20/C23 TT and C24 Tet/C26
TT ratios but the lowest C23 TT/C30 H ratios (Table 1, Figs. 7, 9F
I). Such a biomarker association suggests signicant contribution
from terrigenous organic matter, which is consistent with the
dysoxic conditions that prevailed during the deposition of the
Dongying Formation, as reected by the high Pr/Ph ratios and
low C35 22S/C34 22S hopane, gammacerane/ab C30 hopane and
ETR (Table 1, Fig. 9AD). The C19/C23 TT, C20/C23 TT and C24 Tet/C26
TT ratios for Es3 are slightly higher than those for Es1 but signicantly lower than those for the Dongying Formation (Table 1,
Figs. 7, 9FH), suggesting a minor contribution of terrigenous
organic matter (Hao et al., 2009a,b).
Sterane parameters have the potential to reect primary producers in marine and lacustrine systems (e.g., Volkman et al.,
1998; Knoll et al., 2007; Seplveda et al., 2009). 4-Methylsteranes
are commonly found in marine, evaporitic and especially freshwater environments (e.g., Brassell et al., 1986; Summons et al., 1992;
Peters et al., 2005, p. 530532). The precursors of 4-methylsteranes

are presumed to be 4-methyl sterols. Apart from compounds specically identied as originating from methane oxidizing bacteria
(Jahnke et al., 1999), 4-methyl steroids are mostly algal in origin
(e.g. de Leeuw et al., 1983), and dinoagellates are believed to be
their main source (De Leeuw et al., 1983; Summons et al., 1987).
In the Bohai Bay Basin, abundant 4-methylsteranes appear to be
associated with abundant dinoagellates Bohaidina and Parabohaidina (e.g., Chen et al., 1996, 1998; Zhang et al., 2005). The 4methylsterane index (4-methylsteranes/RC29 steranes) exhibits a
general decrease up section (Fig. 9J), suggesting decreasing contributions from dinoagellates Bohaidina and Parabohaidina. The fact
that Es3 has higher 4-methylsterane abundances than Es1 (Figs. 7
and 9J) supports our interpretation that freshwater lakes
dominated during Es3 deposition whereas saline lakes dominated
during Es1 deposition, since dinoagellates Bohaidina and
Parabohaidina thrive in freshwater settings (Fu et al., 1990; Peters
et al., 2005, p. 531).
It is usually believed that C27 steranes derive mainly from phytoplankton and metazoa, whereas C29 steranes mainly originate
from terrigenous higher plants (e.g., Huang and Meinschein,
1979; Volkman, 1986). C27/C29 sterane ratios range between 0.42
and 0.99 (average 0.76) for Es3 samples, increase to the highest values in the Bozhong sub-basin for Es1 samples (0.821.41, average
0.98), and then decrease to low values (0.481.0, average 0.75)
for the Dongying samples (Fig. 9K). As discussed earlier, C19/C23
TT, C20/C23 TT and C24 Tet/C26 TT are effective parameters reecting
organic matter input from terrigenous higher plants. The Es3 and
Ed have quite similar C27/C29 sterane ratios (Figs. 7, 9K), which appears to be inconsistent with the fact that the Dongying Formation
has much higher C19/C23 TT, C20/C23 TT and C24 Tet/C26 TT ratios
(Table 1, Figs. 7, 9FH) and therefore has signicant contribution
from terrigenous higher plants whereas Es3 was dominated by algal organic matter. The relative C29 sterane abundances show a
general increase with increasing C19/C23 TT and C24 Tet/C26 TT ratios (Fig. 10H, I). The Es3 and Ed have quite different C19/C23 TT
(0.241.01 and 0.591.90 for Es3 and Ed, respectively) and C24
Tet/C26 TT (0.092.74 and 1.366.24 for Es3 and Ed, respectively)
ratios. Yet they have quite similar relative abundances of C29 steranes (36.1450.85% and 36.6251.89% for Es3 and Ed, respectively,
Figs. 7, 10H and I). The elevated C29 sterane abundances for the Es3
that has been conrmed to have minor terrigenous organic matter
input indicate that an additional source for C29 steranes existed.
Volkman et al. (1999) and Kodner et al. (2008) conrmed that
freshwater microalgae may be an important source for C29 steranes. Since freshwater lakes dominated during Es3 deposition, a
plausible explanation for the enhanced C29 sterane abundances relative to the low C19/C23 TT and C24 Tet/C26 TT ratios in Es3 is the
contribution of freshwater microalgae.
C28 steranes are associated with specic phytoplankton types
(e.g., diatoms, Grantham and Wakeeld, 1988; Volkman et al.,
1998) that contain chlorophyll-c (Knoll et al., 2007). In the Bozhong
sub-basin, an overall trend of increasing C28/C29 sterane ratio with
increasing C27/C29 sterane ratio (Fig. 10J) was observed. C28/C29
sterane ratios show moderate values for Es3 (0.490.82, average
0.62), increase to relatively high values between 0.42 and 0.92
(average 0.77) for Es1 and then decrease to low values for the
Dongying Formation (0.300.84, average 0.51) (Table 1, Fig. 9L).
This suggests an enhanced contribution from chlorophyll-c
containing phytoplankton relative to C29 producing organisms
(Knoll et al., 2007) in Es1.
The sterane/hopane ratios reect input of eukaryotic (mainly algae and higher plants) versus prokaryotic (bacteria) organisms to
the source rocks (e.g., Peters and Moldowan, 1993; Gonalves,
2002; Peters et al., 2005, p. 524; Seplveda et al., 2009). In the
Bozhong sub-basin, sterane/hopane ratios increase as C28 sterane
abundances increase and decrease as C29 sterane abundances

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

increase (Fig. 10K and L), indicating that high sterane/hopane


ratios in the Bozhong sub-basin were caused by contribution from
chlorophyll-c containing phytoplankton rather than C29 producing
organisms (Knoll et al., 2007). Sterane/hopane ratios show low values for Es3 (0.100.37, average 0.16), increase to relatively high
values for Es1 (0.160.57, average 0.31), and drop to extremely
low values for the Dongying Formation (0.060.24, average 0.11)
(Table 1, Fig. 9M). The sterane/hopane ratios for Es3 are comparable
with those for the organic-rich lacustrine source rocks dominated
by algal organic matter in the Camamu Basin (Gonalves, 2002)
and in the Congo Basin (Harris et al., 2004). These sterane/hopane
ratios, combined with low C19/C23 TT, C20/C23 TT and C24 Tet/C26 TT
ratios, suggest important contribution from bacterial biomass (e.g.,
Gonalves, 2002; Peters et al., 2005, p. 524; Seplveda et al., 2009).
The relatively high sterane/hopane ratios for Es1 co-occur with relatively high C23 TT/ab C30 hopane ratios and high C27 and C28 sterane abundances (Table 1, Figs. 9 and 10K), and therefore indicate
reduced contribution from bacteria but enhanced primary production rates (Gonalves, 2002). The extremely low sterane/hopane ratios for the Dongying Formation are in accordance with the high
C19/C23 TT, C20/C23 TT and C24 Tet/C26 TT ratios and reect a significant contribution from terrigenous organic matter and heavy bacterial degradation (e.g., Tissot and Welte, 1984, p. 122; Frimmel
et al., 2004; Peters et al., 2005, p. 524).
5.4. Models for source rock deposition under different tectonic and
climatic conditions
Stratigraphic and subsidence studies show that the Bozhong
sub-basin experienced rapid subsidence during Es3 deposition
(up to 600 m/Ma), reduced subsidence rate during Es2 and Es1
deposition (no higher than 300 m/Ma) and another episode of rapid subsidence during Ed deposition (up to 600 m/Ma). The paleoclimate also changed considerably, from humid climate during Es3
deposition through arid climate during Es2 and Es1 deposition to
humid climate again during the deposition of the Dongying Formation (Wang et al., 2010). It is evident that the changes in water
chemistry indicated by biomarker parameters (Fig. 9AE) were
consistent with changes in subsidence rate and climates. More
importantly, both parameters reecting terrigenous higher plant
input (Fig. 9FI) and parameters associated with primary producers (Fig. 9JM) co-vary with parameters reecting depositional
conditions (Fig. 9AE), which strongly suggest the co-evolution of
ecological systems with environments. In other words, changes
in tectonic subsidence and climate induced environmental
changes, which in return resulted in changes in ecological communities in the lake systems. These observed environmental and ecological changes enable us to construct models for the deposition of
high quality lacustrine source rocks under different tectonic and
climatic conditions.
During Es3 deposition, the intense activity of border faults and
rapid subsidence, together with high water inow under wet climate but relatively low sediment input, resulted in a deep, freshwater lake and generally regressional or aggradational basin ll
(Fig. 11A). The low area/depth ratio of the lake probably hampered
efcient water mixing, favoring stable water column stratication
(Gonalves, 2002). The stable water column stratication can be
inferred from the moderate gammacerane indices for the Es3 samples (Fig. 9C) deposited in freshwater lake (e.g., Sinninghe Damst
et al., 1995). The water column stratication resulted probably
from a temperature gradient and a moderate-depth thermocline
was expected (Bohacs et al., 2000), forming a thick and relatively
permanent anoxic water column (Fig. 11A). The oxygen depletion
and the low sulfate availability resulted in the dominance of bacterial fermentation and methanogenesis in the water column/
sediments under the thermocline which was evidenced by the

335

abnormally high abundances of phytane (cf. Fig. 6) believed to be


derived from methanogenic archaea (e.g., Rowland, 1990;
Fuhrmann et al., 2004; see the earlier section). Methanogenesis
in the euxinic bottom water and/or sediments must have sustained
methanotrophic bacteria at the interface between oxic and anoxic
zones (Collister and Wavrek, 1996; Fig. 11A), which can be evidenced from the occurrence of the large isotopic difference between saturate and aromatic fractions for Es3 samples with high
hydrogen indices (Guo, 2009, personal communication). The high
water inow, which carried dissolved inorganic carbon and nitrate
into the lake, might have sustained moderate to high primary productivity. In the freshwater lake, dinoagellates Bohaidina and
Parabohaidina thrived as evidenced from the high 4-methylsterane
indices (Fig. 9J), and freshwater microalgae and chlorophyll-c containing phytoplankton might also be important members of the
community. The contribution of freshwater microalgae caused
the enhanced C29 sterane abundances (Figs. 9K, 10H, I) in the absence of signicant terrigenous higher plant input. The moderate
to high primary productivity and stable water column stratication
accounted for the deposition of the laminated, high quality source
rock in Es3, the most important source rock in the Bozhong subbasin (Gong, 1997; Hao et al., 2009c).
During Es1 deposition, the low water inow and intensive evaporation in arid climate, together with the low subsidence rate, led
to the formation of saline, alkaline, shallower lake (Fig. 11B). Low
sediment input resulted in regressional or aggradational basin-ll,
and culminated in carbonate deposition (Fig. 11B). The weak disturbance by water inow and high salinity favored water column
stratication and the establishment of a shallow, stable chemocline (Fig. 11B). The stable water column stratication and euxinic
bottom water conditions can be evidenced from the fact that Es1
has the lowest Pr/Ph, the highest C35 22S/C34 22S hopane, highest
ETR and especially the highest gammacerane indices (Fig. 9AD)
in the Bozhong sub-basin. The stable water column stratication
and euxinic bottom water conditions can also be inferred from
the high Ph/n-C18 ratios (cf. Fig. 6) and the large isotopic difference
between the saturate and aromatic fractions for the Es1 samples
(Guo, 2009, personal communication), which are believed to be
associated with methanogenic archaea (Rowland, 1990; Fuhrmann
et al., 2004) and methanotrophic bacteria (Collister and Wavrek,
1996), respectively. The combination of relatively high nutrient
concentration in the saline waters (Fuhrmann et al., 2004), abundant CO3 ions under alkaline conditions (Fuhrmann et al., 2004)
and the excellent water column illumination due to very low sediment input (Kelts, 1988) might lead to high primary productivity
in the saline lake (Fig. 11B), which appears to be supported by the
fact that the Es1 has the highest sterane/hopane ratios in the
Bozhong sub-basin (Fig. 9M). In the warm, saline lake, chlorophyll-c containing phytoplankton bloomed, as indicated by the relatively high C28/C29 sterane ratios (Table 1, Fig. 9L). Dinoagellates
Bohaidina and Parabohaidina might also be important members of
phytoplankton community, as indicated by the moderate 4methylsterane indices (Fig. 9J). This scenario accounts for the
deposition of high quality source rock in Es1, which was conrmed
to have independently sourced commercial oil accumulations (Hao
et al., 2010) and made important contribution to many commercial
accumulations in the Bozhong sub-basin (Hao et al., 2009a,b). Due
to smaller thickness and lower thermal maturity, Es1 is of lesser
importance as source rock than Es3 in the Bozhong sub-basin.
As a consequence of rapid subsidence and high water inow,
deep, freshwater lakes occurred during the deposition of the
Oligocene Dongying Formation (Fig. 11C). The lakes at this stage
seemed to be hydrologically open. This, together with the rapid
propagation of delta due to efcient sediment supply, led to a deep,
unstable thermocline with dysoxic bottom water conditions
(Fig. 11C). The unstable water column stratication can be inferred

336

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

High Water Inflow and


Moderate Sediment Supply

(A)
Moderate to High Productivity

TOM

Freshwater, Oxic

TOM

MB

Ciliates

Fresh to brackish, anoxic,


Organic Facies
BC and C

CH4

Organic Facies A and B

High water inflow and moderate sediment supply, regressional or aggradational


basin-fill.
Moderate-depth, stable thermocline, euxinic bottom conditions (methanogenesis).
Low to moderate terrigenous organic matter input, moderate to high primary
productivity, thriving of dinoflagellates Bohaidina and Parabohaidina.

(B)

Low Water Inflow


and Sediment Supply

evaporation
High Productivity

Saline
water,
oxic

Saline, anoxic

Carbonate

CH4
Organic Facies
BC and C
Organic Facies A and B

Low water inflow and sediment supply, intensive evaporation, abundant carbonate
deposition, regressional or aggradational basin-fill.
Shallow, stable chemocline, euxinic bottom conditions (methanogenesis).
Very low terrigenous organic matter input, high primary productivity, blooming of
chlorophyll-c containing phytoplankton.
High Water Inflow and
Efficient Sediment Supply

(C)
Moderate to High Productivity?

TOM

Freshwater, oxic

Dysoxic to anoxic
Organic Facies BC and C
Organic Facies
B + BC

High water inflow and efficient sediment supply, progradational basin-fill.


Deep, unstable thermocline.
High terrigenous organic matter input, algal organic matter preferentially oxidized.
Fig. 11. Depositional models for the three source rock intervals showing the environmental and ecological changes induced by changes in tectonic subsidence and climate
and their control on organic facies distributions. (A) Deep, freshwater lake at Es3 deposition stage; (B) Shallower, saline-alkaline lake at Es1 deposition stage; (C) Freshwater
lake at Ed deposition stage. TOM = terrigenous organic matter; MB = Methanotrophic bacteria. See text for further discussion.

from the very low gammacerane indices (Fig. 9C). The dysoxic
bottom water conditions are indicated by the relatively high Pr/Ph,
low C35 22S/C34 22S hopane and low ETR (Fig. 9B, D), and can

also be inferred from the absence of high Ph/n-C18 ratios (cf.


Fig. 6) which indicated that methanogenesis in the bottom
water/sediments did not occur (Fig. 11C). High water inow, high

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

sediment input and abundant terrigenous organic matter input


might provide efcient nitrate to support moderate to high primary productivity (Harris et al., 2004). However, algal organic matter might be preferentially degraded or oxidized in the oxic or
dysoxic conditions, which appears to be supported by the extremely low sterane/hopane ratios observed in the Ed samples
(Fig. 9M). The oxidation/degradation of algal organic matter, together with dilution by terrigenous organic matter (Sachsenhofer
et al., 2003), led to the deposition of the Dongying Formation
source rock with signicant higher plant organic matter input
(Fig. 11C). Source rock in the Dongying Formation has been conrmed to have independently sourced a small commercial oil accumulation close to the generative kitchens (Hao et al., 2010), and
made important contribution to the Penglai 19-3 oil eld, the largest offshore oil eld so far found in China (Hao et al., 2009c).
In summary, changes in tectonic subsidence and climate during
the syn-rift evolution of the Bozhong sub-basin caused changes in
hydrological status and water chemistry of the lakes, which
induced changes in the phytoplankton community and primary
productivity. The synergetic evolution of environments and organisms in the lake systems (Fig. 11) accounted for the deposition of
the three source rock intervals with different hydrocarbon generating potentials (Figs. 4 and 5) and distinctly different biomarker
assemblages (Figs. 610).

6. Conclusions
Based on our bulk geochemical study of more than 300 samples
and molecular geochemical observation on 57 samples, the following conclusions can be drawn.
(1) The third (Es3) and rst (Es1) members of the Eocene Shahejie
Formation and the Oligocene Dongying Formation (Ed) all
display widely variable total organic carbon contents,
RockEval hydrogen indices, woody and amorphous organic
matter contents, Pr/Ph and Ph/n-C18 ratios, suggesting strong
lateral heterogeneities in organic facies from deep water to
marginal lake sediments. Amorphous organic matter, which
is usually believed to originate mainly from algal organic
matter in reducing environments, may be complex in origin
and therefore may be either rich or poor in hydrogen.
(2) Despite the strong lateral heterogeneities indicated by bulk
geochemical analyses, the deep-lake facies samples from
any interval display fairly narrow variation ranges for most
molecular parameters reecting organic matter origins
and/or depositional environments. This implies that effective oil source rocks (high quality source rocks that have
generated and expelled a considerable amount of oil and
made a considerable contribution to commercial oil
accumulations) in different intervals might not be as
heterogeneous as that reected by the wide variation in
TOC contents, visual kerogen composition and hydrogen
indices, which is particularly important for oil-source rock
correlation. The deep-lake facies from the three intervals,
although all are dominated by shales/mudstones, have
distinctly different biomarker associations, suggesting that
molecular geochemical analysis is a powerful tool to construct the basin ll history.
(3) The distinctly different biomarker assemblages for the three
syn-rift intervals indicate weakly alkaline, freshwater lakes
with moderately deep thermocline during Es3 deposition,
alkaline-saline lakes with shallow chemocline during Es1
deposition and acidic, freshwater lakes with deep, unstable
thermocline during the deposition of the Dongying Formation. Such environmental changes corresponded to changes

337

in subsidence rate and paleoclimate, from rapid subsidence


and wet climate during Es3 deposition, through slow subsidence and arid climate during Es1 deposition to rapid subsidence and wet climate during Ed deposition and resulted in
synchronous changes in terrigenous organic matter input,
phytoplankton community and primary productivity. The
synergetic evolution of environments and organisms controlled by tectonic subsidence and climate accounted for
the deposition and distribution of high quality lacustrine
source rocks with distinctly different geochemical characteristics. The development of source rocks at three syn-rift
stages under different tectonic and climatic conditions is
responsible for the great petroleum reserves and the diverse
geochemical characteristics of petroleum in the sub-basin.
Acknowledgments
This study was supported by the National Natural Science
Foundation of China (Grant No. 90914006) and Program for
Changjiang Scholars and Innovative Research Team in the
University (IRT0658). We appreciate the collaboration and enthusiastic support of Qinglong Xia, Lixin Tian, Shui Yu, Yonghua Guo and
Shike Zhou at the Tianjin Branch of China National Offshore Oil
Company Ltd, Zaisheng Gong, Yunhua Deng, Jingfu Wu, Yimei
Sun and Gongcheng Zhang at the CNOOC Research Center, Xiongqi
Pang, Chunjiang Wang, Xiuxiang L, Liangjie Tang, Jiafu Qi,
Nansheng Qiu, Zhenxue Jiang and many other co-workers at China
University of Petroleum, Sitian Li, Renchen Xin and Qiugen Zheng
at China University of Geosciences at Beijing. We thank Tianjin
Branch of China National Offshore Oil Company Ltd. for kindly providing part of the data in this study. L.R. Snowdon (Editor-in-Chief)
is gratefully acknowledged for his constructive review of an earlier
version of this manuscript. We thank Barry J. Katz, Maowen Li and
Kenneth E. Peters (Associate Editor) for their thorough and critical
reviews and suggestions to improve the manuscript.
Associate EditorKenneth E. Peters
References
Allen, M.B., McDonald, D.I.M., Zhao, X., Vincent, S.J., Brouet-Menzies, C., 1997. Early
Cenozoic two-phase extension and late Cenozoic thermal subsidence and
inversion of the Bohai basin, northern China. Marine and Petroleum Geology 14,
951972.
Allen, M.B., McDonald, D.I.M., Zhao, X., Vincent, S.J., Brouet-Menzies, C., 1998.
Transtensional deformation in the evolution of the Bohai basin, northern China.
In: Holdsworth, R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental Transpression
and Transtensional Tectonics. Geological Society (London) Special Publication
135, 215229.
Bennett, B., Olsen, S.D., 2007. The inuence of source depositional conditions on the
hydrocarbon and nitrogen compounds in petroleum from central Montana,
USA. Organic Geochemistry 38, 935956.
Bohacs, K.M., Carroll, A.R., Neal, J.E., Mankiewicz, P.J., 2000. Lake-basin type, source
potential, and hydrocarbon character. An integrated sequence-stratigraphicgeochemical framework. In: Gierlowski-Kordesch, E.H., Kelts, K.R. (Eds.), Lake
Basins through Space and Time. American Association of Petroleum Geologists
Studies in Geology 46, 334.
Brassell, S.C., Eglinton, G., Fu, J.M., 1986. Biological marker compounds as indicators
of the depositional history of the Maoming oil shale. In: Leythaeuser, D.,
Rullktter, J. (Eds.), Advances in Organic Geochemistry 1985. Pergamon Press,
Oxford, pp. 927941.
Brincat, D., Abbott, G., 2001. Some aspects of the molecular biogeochemistry of
laminated and massive rocks from the Naples Beach Section (Santa Barbara
Ventura Basin). In: Isaacs, C.M., Rullktter, J. (Eds.), The Monterey Formation:
From Rocks to Molecules. Columbia University Press, New York, pp. 140169.
Carroll, A.R., Bohacs, K.M., 1999. Stratigraphic classication of ancient lakes:
balancing tectonic and climatic controls. Geology 27, 99102.
Carroll, A.R., Bohacs, K.M., 2001. Lake-type controls on petroleum source rock
potential in nonmarine basins. American Association of Petroleum Geologists
Bulletin 85, 10331053.
Chang, C.Y., 1991. Geological characteristics and distribution patterns of
hydrocarbon deposits in the Bohai Bay basin, east China. Marine and
Petroleum Geology 8, 98106.

338

F. Hao et al. / Organic Geochemistry 42 (2011) 323339

Chen, J.Y., Bi, Y.P., Zhang, J.G., Li, S.F., 1996. Oil-source correlation in the Fulin basin,
Shengli petroleum province, East China. Organic Geochemistry 24, 931940.
Chen, J.Y., Li, S.F., Xiong, Y., Bi, Y.P., 1998. Multiple petroleum systems in Tertiary
extentional basins, East China: a case study of the GunanFulin basin. Journal of
Petroleum Geology 21, 105118.
Collister, J.W., Wavrek, D.A., 1996. D13C compositions of saturate and aromatic
fractions of lacustrine oils and bitumens: evidence for water column
stratication. Organic Geochemistry 24, 913920.
Curiale, J.A., 2008. Oil-source rock correlations limitations and recommendations.
Organic Geochemistry 39, 11501161.
De Grande, S.M.B., Aquino Neto, F.R., Mello, M.R., 1993. Extended tricyclic terpanes
in sediments and petroleums. Organic Geochemistry 20, 10391047.
De Leeuw, J.W., Rijpstra, W.I.C., Schenck, P.A., Volkman, J.K., 1983. Free, esteried
and residual bound sterols in Black Sea Unit I sediments. Geochimica et
Cosmochimica Acta 47, 455465.
Demaison, G.J., Moore, G.T., 1980. Anoxic environments and oil source bed genesis.
American Association of Petroleum Geologists Bulletin 64, 11791209.
Ebukanson, E.J., Kinghorn, R.R.F., 1985. Kerogen facies in the major Jurassic
mudrock formations of southern England and the implication on the
depositional environments of their precursors. Journal of Petroleum Geology
8, 435462.
Ercegovac, M., Kostic, A., 2006. Organic facies and palynofacies: nomenclature,
classication and applicability for petroleum source rock evaluation. International Journal of Coal Geology 68, 7078.
Fan, M.J., Dettman, D.L., Song, C.H., Fang, X.M., Garzione, C.N., 2007. Climatic
variation in the Linxia basin, NE Tibetan Plateau, from 13.1 to 4.3 Ma: the stable
isotope record. Palaeogeography, Palaeoclimatology, Palaeoecology 247, 313
328.
Follows, B., Tyson, R.V., 1998. Organic facies of the Asbian (early Carboniferous)
Queensferry Beds, Lower Oil Shale Group, South Queensferry, Scotland, and a
brief comparison with other Carboniferous North Atlantic oil shale deposits.
Organic Geochemistry 29, 821844.
Frimmel, A., Oschmann, W., Schwark, L., 2004. Chemostratigraphy of the Posidonia
Black Shale, SW Germany I. Inuence of seal-level variation on organic facies
evolution. Chemical Geology 206, 199230.
Fu, J.M., Sheng, G., Xu, J., Eglinton, G., Gowar, A.P., Jia, R., Fan, S., 1990. Application of
biological markers in the assessment of paleoenvironments of Chinese nonmarine sediments. Organic Geochemistry 16, 769779.
Fuhrmann, A., Horseld, B., Lpez, J.F., Hu, L.G., Zhang, Z.W., 2004. Organic facies,
depositional environment and petroleum generating characteristics of the
lacustrine Shahejie Formation, Es4 Member, western Depression, Liaohe Basin
(NE China). Journal of Petroleum Geology 27, 2746.
Ge, T., Chen, Y., 1993. Liaohe Oileld. Petroleum Geology of China III. Petroleum
Industry Press, Beijing. 237 p. (in Chinese).
George, S.C., Lisk, M., Eadington, P.J., 2004. Fluid inclusion evidence for an early,
marine-sourced oil charge prior to gas-condensate migration, Bayu-1, Timor
Sea, Australia. Marine and Petroleum Geology 21, 11071128.
Gonalves, F.T.T., 2002. Organic and isotope geochemistry of the Early Cretaceous
rift sequence in the Camamu Basin, Brazil: paleolimnological inferences and
source rock models. Organic Geochemistry 33, 6780.
Gong, Z.S., 1997. Giant Offshore Oil and Gas Fields in China. Petroleum Industry
Press, Beijing. 396 p. (in Chinese).
Gong, Z.S., Zhu, W.L., Chen, P.P.H., 2010. Revitalization of a mature oil-bearing basin
by a paradigm shift in the exploration concept. A case history of Bohai Bay,
Offshore China. Marine and Petroleum Geology 27, 10111027.
Grantham, P.J., Wakeeld, L.L., 1988. Variations in the sterane carbon number
distributions of marine source rock derived crude oils through geological times.
Organic Geochemistry 12, 6177.
Hanson, A.D., Zhang, S.C., Moldowan, J.M., Liang, D.G., Zhang, B.M., 2000. Molecular
organic geochemistry of the Tarim Basin, Northwest China. American
Association of Petroleum Geologists Bulletin 84, 11091128.
Hanson, A.D., Ritts, B.D., Zinniker, D., Moldowan, J.M., Bif, U., 2001. Upper
Oligocene lacustrine source rocks and petroleum systems of the northern
Qaidam basin, northwest China. American Association of Petroleum Geologists
Bulletin 85, 601619.
Hao, F., Chen, J.Y., Sun, Y.C., Liu, Y.Z., 1993. Application of organic facies studies to
sedimentary basin analysisa case study from the Yitong graben, China.
Organic Geochemistry 20, 2743.
Hao, F., Zou, H.Y., Gong, Z.S., Deng, Y.H., 2007. Petroleum migration and
accumulation in the Bozhong sub-basin, Bohai Bay basin, China: signicance
of preferential petroleum migration pathways (PPMP) for the formation of large
oilelds in lacustrine fault basins. Marine and Petroleum Geology 24,
113.
Hao, F., Zhou, X.H., Zhu, Y.M., Yang, Y.Y., 2009a. Mechanisms for oil depletion and
enrichment on the Shijiutuo uplift, Bohai Bay basin, China. American
Association of Petroleum Geologists Bulletin 93, 10151037.
Hao, F., Zhou, X.H., Zhu, Y.M., Zou, H.Y., Bao, X.H., Kong, Q.Y., 2009b. Mechanisms of
petroleum accumulation in the Bozhong sub-basin, Bohai Bay basin, China. Part
1: origin and occurrence of crude oils. Marine and Petroleum Geology 26, 1528
1542.
Hao, F., Zhou, X.H., Zhu, Y.M., Bao, X.H., Yang, Y.Y., 2009c. Charging of the Neogene
Penglai 193 eld, Bohai Bay basin, China: oil accumulation in a young trap in
an active fault zone. American Association of Petroleum Geologists Bulletin 93,
155179.
Hao, F., Zhou, X.H., Zhu, Y.M., Zou, H.Y., Yang, Y.Y., 2010. Charging of oil elds
surrounding the Shaleitian Uplift from multiple source rock intervals and

generative kitchens, Bohai Bay basin, China. Marine and Petroleum Geology 27,
19101926.
Hao, F., Zhang, Z.H., Zou, H.Y., Zhang, Y.C., Yang, Y.Y., in press. Origin and mechanism
of formation of the low-oil-saturation Moxizhuang eld, Junggar basin, China:
implication for petroleum exploration in basins having complex histories.
American Association of Petroleum Geologists Bulletin.
Harris, N.B., Freeman, K.H., Pancost, R.D., White, T.S., Mitchell, G.D., 2004. The
character and origin of lacustrine source rocks in the Lower Cretaceous synrift
section, Congo Basin, west Africa. American Association of Petroleum Geologists
Bulletin 88, 11631184.
Holba, A.G., Ellis, L., Dzou, I.L., 2001. Extended tricyclic terpanes as age discriminators between Triassic, Early Jurassic and Middle-Late Jurassic oils. Presented
at the 20th International Meeting on Organic Geochemistry, 1014 September,
2001, Nancy, France.
Holba, A.G., Dzou, L.I., Wood, G.D., Ellis, L., Adam, P., Schaeffer, P., Albrecht, P., Greene,
T., Hughes, W.B., 2003. Application of tetracyclic polyprenoids as indicators of
input from fresh-brackish water environments. Organic Geochemistry 34,
441469.
Hsiao, L.Y., Graham, S.A., Tilander, N., 2004. Seismic reection imaging of a major
strike-slip fault zone in a rift system: paleogene structure and evolution of the
Tan-Lu fault system, Liaodong Bay, Bohai, offshore China. American Association
of Petroleum Geologists Bulletin 88, 7197.
Hsiao, L.Y., Graham, S.A., Tilander, N., 2010. Stratigraphy and sedimentation in a rift
basin modied by synchronous strike-slip deformation: southern Xialiao basin,
Bohai, offshore China. Basin Research 22, 6178.
Huang, W.Y., Meinschein, W.G., 1979. Sterols as ecological indicators. Geochimica et
Cosmochimica Acta 43, 739745.
Huc, A.Y., Lallier-Verges, E., Bertrand, P., Carpentier, B., Hollander, D.J., 1992. Organic
matter response to change of depositional environment in Kimmeridgian
Shales, Dorset, UK. In: Whelan, J.K., Farrington, J.W. (Eds.), Organic Matter:
Productivity. Accumulation, and Preservation in Recent and Ancient Sediments.
Columbia University Press, New York, pp. 469486.
Jahnke, L.L., Summons, R.E., Hope, J.M., Des Marais, D.J., 1999. Carbon isotopic
fractionation in lipids from methanotrophic bacteria II: the effects of physiology
and environmental parameters on the biosynthesis and isotopic signatures of
biomarkers. Geochimica et Cosmochimica Acta 63, 7993.
Jones, R.W., 1987. Organic facies. In: Brooks, J., Welte, D. (Eds.), Advances in
Petroleum Geochemistry II, pp. 190.
Jones, F.G., Bowser, C.J., 1978. The mineralogy and related chemistry of lake sediments.
In: Lerman, A. (Ed.), Lakes: Chemistry, Geology and Physics. Springer, pp. 179
235.
Justwan, H., Dahl, B., Isaksen, G.H., 2006. Geochemical characterization and genetic
origin of oils and condensates in the South Viking Graben, Norway. Marine and
Petroleum Geology 23, 213239.
Katz, B.J., 1990. Controls on distribution of lacustrine source rocks through time and
space. In: Katz, B.J. (Ed.), Lacustrine Basin Exploration: Case Studies and Modern
Analogs. American Association of Petroleum Geologists Memoir 50, 6176.
Katz, B.J., 1995. A survey of rift basin source rocks. In: Lambiase, J.J. (Ed.),
Hydrocarbon Habitat of Rift Basins. Geological Society (London) Special
Publication 80, 213242.
Katz, B.J., 2005. Controlling factors on source rock developmenta review of
productivity, preservation and sedimentation rate. In: Harris, N.B. (Ed.), The
deposition of organic-carbon-rich sediments: models, mechanisms, and
consequences. SEPM Special Publication 82, 716.
Keely, B.J., Sinninghe Damst, J.S., Betts, S.E., Yue, L., de Leeuw, J.W., Maxwell, J.R.,
1993. A molecular stratigraphic approach to palaeoenvironmental assessment
and the recognition of changes in source inputs in marls of the Mulhouse Basin
(Alsace, France). Organic Geochemistry 20, 11651186.
Kelts, K., 1988. Environments of deposition of lacustrine petroleum source rocks: an
introduction. In: Fleet, A.J., Kelts, K., Talbot, M.R. (Eds.), Lacustrine Petroleum
Source Rocks. Geological Society Special Publication 40, 326.
Keym, M., Dieckmann, V., Horseld, B., Erdmann, M., Galimberti, R., Kua, L.C., Leith,
L., Podlaha, O., 2006. Source rock heterogeneity of the Upper Jurassic Draupne
Formation, north Viking Graben, and its relevance to petroleum generation
studies. Organic Geochemistry 37, 220243.
Kirkland, D.W., Evans, R., 1981. Source-rock potential of evaporitic environment.
American Association of Petroleum Geologists Bulletin 65, 181190.
Knoll, A.H., Summons, R.E., Waldbauer, J.R., Zumberge, J.E., 2007. The geological
succession of primary producers in the oceans. In: Falkowski, P., Knoll, A.H.
(Eds.), The Evolution of Primary Producers in the Sea. Academic Press, Boston,
pp. 133163.
Kodner, R.B., Summons, R.E., Pearson, A., King, N., Knoll, A.H., 2008. Sterols in a
unicellular relative of the metazoans. Proceedings of the National Academy of
Sciences of the USA 105, 98979902.
Koopmans, M.P., de Leeuw, J.W., Lewan, M.D., Sinninghe Damst, J.S., 1996. Impact
of dia- and catagenesis on sulfur and oxygen sequestration of biomarkers as
revealed by articial maturation of an immature sedimentary rock. Organic
Geochemistry 25, 391426.
Kruge, M.A., Hubert, J.F., Akes, R.J., Meriney, P.E., 1990a. Biological markers in Lower
Jurassic synrift lacustrine black shales, Hartford basin, Connecticut, USA.
Organic Geochemistry 15, 281289.
Kruge, M.A., Hubert, J.F., Bensley, D.F., Crelling, J.C., Akes, R.J., Meriney, P.E., 1990b.
Organic geochemistry of a Lower Jurassic synrift lacustrine sequence, Hartford
basin, Connecticut, USA. Organic Geochemistry 16, 671689.
Lu, K.Z., Qi, J.F., 1997. Structural Model for the Cenozoic Petroliferous Basins in the
Bohai Bay. Geological Publishing House, Beijing. 207 p. (in Chinese).

F. Hao et al. / Organic Geochemistry 42 (2011) 323339


McKenzie, D., 1978. Some remarks on the development of sedimentary basins. Earth
and Planetary Science Letters 40, 2532.
Meyers, P.A., 1997. Organic geochemical proxies of paleoceanographic,
paleolimnologic, and paleoclimatic processes. Organic Geochemistry 27, 213
250.
Meyers, P.A., 2003. Applications of organic geochemistry to paleolimnological
reconstructions: a summary of examples from the Laurentian Great Lakes.
Organic Geochemistry 34, 261289.
Ohm, S.E., Karlsen, D.A., Austin, T.J.F., 2008. Geochemically driven exploration
models in uplifted area: examples from the Norwegian Barents Sea. American
Association of Petroleum Geologists Bulletin 92, 11911223.
Peters, K.E., 1986. Guidelines for evaluating petroleum source rock using
programmed pyrolysis. American Association of Petroleum Geologists Bulletin
70, 318329.
Peters, K.E., Moldowan, J.M., 1993. The Biomarker Guide: Interpreting Molecular
Fossils in Petroleum and Ancient Sediments. Prentice-Hall, New Jersey. 363
p.
Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide, Biomarkers
and Isotopes in Petroleum Exploration and Earth History. Cambridge University
Press, Cambridge. 1155 p..
Philp, R.P., Gilbert, T.D., 1986. Biomarker distributions in Australian oils
predominantly derived from terrigenous source material. Organic
Geochemistry 10, 7384.
Powell, T.G., 1986. Petroleum geochemistry and depositional setting of lacustrine
source rocks. Marine and Petroleum Geology 3, 200219.
Powell, A.J., Dodge, J.D., Lewis, J., 1990. Late Neogene to Pleistocene palynological
facies of the Peruvian continental margin upwelling, Leg 112. In: Suess, E., Von
Huene, R. (Eds.), Proceeding of the Ocean Drilling Project, Scientic Results, vol.
112. ODP, College Station, Texas, pp. 297321.
Preston, J.C., Edwards, D.S., 2000. The petroleum geochemistry of oils and source
rocks from the northern Bonaparte Basin, offshore northern Australia. The APEA
Journal 40, 257282.
Price, L.C., 1989. Primary petroleum migration from shales with oxygen-rich organic
matter. Journal of Petroleum Geology 12, 289324.
Qi, J.F., Yang, Q., 2010. Cenozoic structural deformation and dynamic processes of
the Bohai Bay basin province, China. Marine and Petroleum Geology 27, 757
771.
Ritts, B.D., Hanson, A.D., Zinniker, D., Moldowan, J.M., 1999. Lower-Middle Jurassic
nonmarine source rocks and petroleum systems of the northern Qaidam basin,
northwest China. American Association of Petroleum Geologists Bulletin 83,
19802005.
Rowland, S.J., 1990. Production of acyclic isoprenoid hydrocarbons by laboratory
maturation of methanogenic bacteria. Organic Geochemistry 15, 916.
Sachsenhofer, R.F., Bechtel, A., Reischenbacher, D., Weiss, A., 2003. Evolution of
lacustrine systems along the Miocene Mur-Mrz fault system (Eastern Alps,
Austria) and implications on source rocks in pull-apart basins. Marine and
Petroleum Geology 20, 83110.
Seplveda, J., Wendler, J., Leider, A., Kuss, H.J., Summons, R.E., Hinrichs, K.U., 2009.
Molecular isotopic evidence of environmental and ecological changes across the
CenomanianTuronian boundary in the Levant Platform of central Jordan.
Organic Geochemistry 40, 553568.
Sinninghe Damst, J.S., Kenig, F., Koopmans, M.P., Koster, J., Schouten, S., Hayes, J.M.,
De Leeum, J.W., 1995. Evidence for gammacerane as an indicator of
water column stratication. Geochimica et Cosmochimica Acta 59, 1895
1900.
Sofer, Z., 1984. Stable carbon isotope compositions of crude oils: application to
source depositional environments and petroleum alteration. American
Association of Petroleum Geologists Bulletin 68, 3149.

339

Summons, R.E., Volkman, J.K., Boreham, C.J., 1987. Dinosterane and other steroidal
hydrocarbons of dinoagellate origin in sediments and petroleum. Geochimica
et Cosmochimica Acta 51, 30753082.
Summons, R.E., Thomas, J., Maxwell, J.R., Boreham, C.J., 1992. Secular and environmental constraints on the occurrence of dinosterane in sediments. Geochimica
et Cosmochimica Acta 56, 24372444.
Summons, R.E., Hope, J.M., Swart, R., Walter, M.R., 2008. Origin of Nama
basin bitumen seeps: petroleum derived from a Permian lacustrine source
rock traversing southwestern Gondwana. Organic Geochemistry 39,
589607.
Talbot, M.R., 1988. The origins of lacustrine source rocks: evidence from
the lakes of tropical Africa. In: Fleet, A.J., Kelts, K., Talbot, M.R. (Eds.),
Lacustrine Petroleum Source Rocks. Geological Society Special Publication 40,
2943.
ten Haven, H.L., Rohmer, M., Rullktter, J., Bisseret, P., 1989. Tetrahymanol, the most
likely precursor of gammacerane, occurs ubiquitously in marine sediments.
Geochimica et Cosmochimica Acta 53, 30733079.
Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence, second ed.
Berlin, Springer-Verlag. 699 p..
Tyson, R.V., 1995. Sedimentary Organic Matter, Organic Facies and Palynofacies.
Chapman and Hall, London. 615 p..
Tyson, R.V., 2005. The productivity versus preservation controversy: cause, aws
and resolution. In: Harris, N.B. (Ed.), The Deposition of Organic-carbon-rich
Sediments: Models, Mechanisms, and Consequences. SEPM Special Publication
82, 1733.
Valero Garcs, B.L., Kelts, K., Ito, E., 1995. Oxygen and carbon isotope trends and
sedimentological evolution of a meromitic and saline lacustrine system: the
Holocene Medicine Lake basin, North American Great Plains, USA.
Palaeogeography, Palaeoclimatology, Palaeoecology 117, 253278.
Venkatesan, M.I., 1989. Tetrahymanol: its widespread occurrence and geochemical
signicance. Geochimica et Cosmochimica Acta 53, 30953101.
Volk, H., George, S., Middleton, H., Schoeld, S., 2005. Geochemical comparison of
uid inclusion and present-day oil accumulations in the Papuan Foreland
evidence for previously unrecognised petroleum source rocks. Organic
Geochemistry 36, 2951.
Volkman, J.K., 1986. A review of sterol markers for marine and terrigenous organic
matter. Organic Geochemistry 9, 8399.
Volkman, J.K., Barrett, S.M., Blackburn, S.I., Mansour, M.P., Sikes, E.L., Gelin, F., 1998.
Microalgal biomarkers: a review of recent research developments. Organic
Geochemistry 29, 11631179.
Volkman, J.K., Barrett, S.M., Blackburn, S.I., 1999. Eustigmatophyte microalgae are
potential sources of C29 sterols, C22C28 n-alcohols and C28C32 n-alkyl diols in
freshwater environments. Organic Geochemistry 30, 307318.
Wang, B., Qian, K., 1992. Geology Research and Exploration Practice in the
Shengli Petroleum Province (in Chinese). Petroleum University Press,
Dongying. 319 p..
Wang, G.L., Li, S., Wang, T.G., Zhang, L.Y., 2010. Applications of molecular fossils in
lacustrine stratigraphy. Chinese Journal of Geochemistry 29, 1520.
Wolfbauer, C.A., Surdam, R.C., 1974. Origin of nonmarine dolomite in Eocene Lake
Gosiute, Green River Basin, Wyoming. Bulletin of Geological Society of America
85, 17331740.
Yang, Y.T., Xu, T.G., 2004. Hydrocarbon habitat of the offshore Bohai basin, China.
Marine and Petroleum Geology 21, 691708.
Ye, H., Shedlock, K.M., Hellinger, S.J., Sclater, J.G., 1985. The North China basin: an
example of a Cenozoic rifted intraplate basin. Tectonics 4, 153169.
Zhang, L.Y., Liu, Q., Zhang, C.R., 2005. Study on the Genetic Relationships between
Hydrocarbon Occurrence and Pool Formation in the Dongying Depression.
Beijing, Geological Publishing House. 202 p. (in Chinese).

You might also like