Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

Journal Pre-proof

Organic matter enrichment and hydrocarbon accumulation models


of the marlstone in the Shulu Sag, Bohai Bay Basin, Northern
China

Xiangxin Kong, Zaixing Jiang, Chao Han, Ruifeng Zhang

PII: S0166-5162(19)30372-6
DOI: https://doi.org/10.1016/j.coal.2019.103350
Reference: COGEL 103350

To appear in: International Journal of Coal Geology

Received date: 2 April 2019


Revised date: 10 November 2019
Accepted date: 19 November 2019

Please cite this article as: X. Kong, Z. Jiang, C. Han, et al., Organic matter enrichment and
hydrocarbon accumulation models of the marlstone in the Shulu Sag, Bohai Bay Basin,
Northern China, International Journal of Coal Geology(2018), https://doi.org/10.1016/
j.coal.2019.103350

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2018 Published by Elsevier.


Journal Pre-proof

Organic matter enrichment and hydrocarbon accumulation models of the

marlstone in the Shulu Sag, Bohai Bay Basin, Northern China

Xiangxin Kong 1,2,3 , Zaixing Jiang 1,3 * , Chao Han2 , Ruifeng Zhang4

1 School of Energy Resources, China University of Geosciences (Beijing), Beijing 100083, China

2 College of Earth Science and Engineering, Shandong University of Science and Technology,

Qindao 266590, China

f
3 Key Laboratory of Marine Reservoir Evolution and Hydrocarbon Enrichment Mechanism,

oo
Ministry of Education, China University of Geosciences (Beijing), Beijing 100083, China

pr
4 PetroChina Huabei Oilfield Company, Renqiu, 062552, China
e-
* Corresponding author. E-mail address: Jiangzx@cugb.edu.cn

ABSTRACT
Pr

Hybrid unconventional systems are composed of juxtaposed organic-rich and organic-lean

intervals. Organic matter enrichment, hydrocarbon migration, and the spatial relationship of
al

source rocks and reservoirs have important influences on differential hydrocarbon accumulation
rn

within the system. Most studies have ignored the complexity of hydrocarbon accumulation in
u

unconventional oil reservoirs. A typical hybrid lithology system with a combination of marlstone
Jo

and rudstone has been discovered in the lower part of the third member of the Shahejie Formation

of the Shulu Sag, Bohai Bay Basin, Northern China. The basin showcases a strong heterogeneity

in unconventional hydrocarbon production and is an appropriate case to study hydrocarbon

accumulation within hybrid unconventional systems. However, relevant work in the study area

was scarce. A study of the mechanisms of organic matter enrichment and hydrocarbon

accumulation was required to understand the heterogeneity of the oil reservoirs. Nine units in

ascending order, namely I L, IU , IIL , IIU , IIIL , IIIU, IV, VL , and VU , were proposed in this study. A

detailed geochemical analysis was carried out to distinguish the organic matter sources and

preservation conditions among the different units, and assess the migration patterns of the

hydrocarbons. The results show that Unit II L has the highest total organic carbon (TOC) content.

1
Journal Pre-proof

Unit II L has a higher quantity of aquatic sourced organic matter, w ith relatively high abundance of

short chain n-alkanes, C27 /C29 sterane, and 4-methyl steranes. Based on the low ratios of pristane

to phytane, the varve-like lamination structure of the lithology and suitable salinity, w e can

speculate that good organic matter preservation conditions existed in Unit II L . Maturity of the

source rocks increases with depth. Biomarkers indicate that hydrocarbons within the marlstones

were retained in situ, whereas hydrocarbons within the rudstones were derived from source rocks

via short- or long-distance migrations. Fractures can act as effective pathways for hydrocarbon

migration. Organic matter sources, preservation conditions, maturity, and hydrocarbon migration

ways jointly controlled the formation of oil reservoirs in the Shulu Sag. The results of this study

f
oo
provide a better understanding of organic matter enrichment and hydrocarbon accumulation and

guide further exploration in the basin.


pr
e-
Keywords: Marlstone; Organic matter; Hydrocarbon migration; Shulu Sag; Biomarker
Pr

1. Introduction
As the demand for unconventional oil and gas resources gradually increases, shale and tight
al

reservoirs have become key exploration target areas in global petroleum geology (Jarvie, 2012a, b;
rn

Han et al., 2016; Liang et al., 2017). A common understanding is that most unconventional sources,

especially shale oil, are stored in source rocks; however, hydrocarbon migration from source rocks
u

into nearby organic-lean intervals is a noteworthy phenomenon in the unconventional system


Jo

(Jarvie et al., 2012b; Kong et al., 2019a). For example, the Bakken shale oil is generated in the

upper and lower mudstone sections and accumulated in the tight carbonate and sandstone

reservoirs within the middle member (Smith and Bustin, 2000). The hydrocarbon migration

process within the unconventional system mainly occurs in a hybrid lithology association, which

is composed of organic-rich mudstone and organic-lean, low-permeability sandstone, rudstone,

and carbonate (Jarvie et al., 2012b). For the hybrid lithology system, the quality of source rocks

controls the potential of unconventional sources (Liang et al., 2017), and organic-lean lithology

reservoirs influence hydrocarbon accumulation and production scale (Lei et al., 2015). Therefore,

organic matter enrichment and hydrocarbon migration within unconventional systems are key

factors to control the formation of hydrocarbon accumulation, such as the Granite Wash tight
2
Journal Pre-proof

rudstone oil reservoir in the Anadarko Basin (Mitchell, 2011), Eagle Ford shale oil reservoir in the

Texas (Jarvie et al., 2012b), and Yanchang tight mudstone-siltstone reservoir in the Ordos Basin

(Xu et al., 2017). The spatial relationship between source rocks and reservoirs has an important

influence on hydrocarbon migration, resulting in differences in unconventional hydrocarbon

accumulation; however, this aspect is easily neglected in research.

Following the successful exploitation of marine unconventional reservoirs in North America

(Smith and Bustin, 2000; Ko et al., 2017), multiple potential unconventional resources have been

found in lacustrine basins in China, such as Bohai Bay Basin (Zhao et al., 2014; Liang et al.,

2018b), Jianghan Bas in (Li et al., 2018), Jiuquan Basin (Guo et al., 2018), Ordos Basins (Li et al.,

f
oo
2016), and Songliao Bas in (Zou et al., 2019). Lacustrine unconventional systems showcase strong

heterogeneity in relation to geochemical and lithological characteristics (Liang et al., 2018a, b; Hu


pr
et al., 2018), reservoir quality (Wang et al., 2017; Kong et al., 2019a), and hydrocarbon migration
e-
(Tang et al., 2018), which limits our understanding of lacustrine unconventional oil and gas

systems, and the prediction of hydrocarbon distribution. The Shulu Sag is a small half-graben
Pr

basin in the Bohai Bay Basin. It has recently become a research hotspot because it is estimated

abundant unconventional resources are present in its huge thickness of lacustrine


al

marlstone-rudstone strata (Zhao et al., 2014). Previous studies suggest that the hydrocarbons in
rn

marlstones are self-sourced, while those in rudstone reservoirs are from nearby organic matter-rich

marlstones (Li et al., 2017). Vertical hydrocarbon migration is proposed based on the
u

heterogeneity of thermal maturity biomarkers (Tang et al., 2018). Moreover, production results
Jo

show that the main oil production is from lower units in the formation (Kong et al., 2017), which

have a relationship with vertical organic and lithological characteristics (Kong et al., 2019a).

Therefore, the Shulu Sag is an appropriate case to study the heterogeneity of continental

unconventional hydrocarbon accumulation within hybrid lithology system. Controlling factors of

organic matter enrichment and hydrocarbon accumulation in the Shulu Sag must be understood;

however, relevant work was scarce in the study area.

The major goals of this work are as follows: (1) to study the mechanism of organic matter

enrichment and discuss the relationship between depositional environments and hydrocarbon

generation potential; and (2) to analyse the migration characteristics of hydrocarbons in the Shulu

Sag and build a model of hydrocarbon accumulation to predict exploration potential in the basin.
3
Journal Pre-proof

2. Geological setting
From the latest Cretaceous through the Paleogene, active mantle convection and the

India-Eurasia collision have caused lithospheric extension and thinning beneath the Bohai Bay

Basin, North China (Ye et al., 1987). This tectonic process created a series of grabens and

half-grabens within the Bohai Bay Basin, such as the Jizhong Depression (Fig. 1). The Shulu Sag,

which lies in the southwestern corner of the Jizhong Depression, is a NE-SW-trending elongated

half-graben basin (Figs. 1, 2A). It formed during the early Paleogene, c ontrolled by the

downthrow of the Xinhe Fault (Jiang et al., 2007). The Shulu Sag is bounded by the Xinhe Fault

f
oo
to the southeast and Ningjin Uplift to the west, and contains three tectonic belts from west to east:

gentle slope, subsag, and steep slope (Fig. 2B). The W-E to WNW-ESE trending faults, including
pr
the Hengshui, Taijiazhuang, and Jingqiu faults, divide the basin into three segments from north to
e-
south (Fig. 2B).

The Paleogene lacustrine strata in the Shulu Sag directly overlapped the
Pr

Cambrian-Ordovician and Permo-Carboniferous marine carbonate strata and gradually onlapped

updip onto the western slope to the Ningjin Uplift (Jiang et al., 2017). In descending order, the
al

overlying strata include the Pingyuan, Minghuazhen, Guantao, Dongying, and Shahejie formations
rn

(Fig. 2C). The Shahejie Formation consist of three members; these are: No. 1 (Es 1 ), No. 2 (Es 2 ),

and No. 3 (Es 3) (Jiang et al., 2017). The Shahejie 3 (Es 3 ) Member can be further divided into
u

upper and lower submembers. Organic matter-rich marlstones containing carbonate rudstones are
Jo

L
developed in the lower submember of Shahejie 3 (Es 3 ), and these are the exploration target for

unconventional sources (Zhao et al., 2014).

A previous study divided Es 3 L submember into five units: I, II, III, IV, and V (Zheng et al.,

2015). The stratigraphic classification of Es 3L in this study is an improved version combined with

the previous scheme and characteristics of organic geochemistry and lithology. Units I, II, III, and

V contain lower and upper units based on changes of lithology and TOC contents. Thick rudstone

intervals are mainly developed in units I L and IIIL. Other units are mainly composed of

marlstones.

3. Sample and methods


4
Journal Pre-proof

Six wells were selected in this study (Fig. 2), including Well J94 (located in the upper gentle

slope zone), Well J97, Well J98X, Well J116X and Well ST1H (located in the lower gentle slope

zone), and Well JG13 (located in the subsag zone adjacent to the steep slope).

Samples for TOC, Rock-Eval pyrolysis, kerogen elements, whole-rock mineralogical

analyses were ground to < 200 mesh using an agate pestle and mortar. The TOC content was

measured by a LECO CS-200 carbon/sulphur instrument after carbonate removal using excessive

diluted hydrochloric acid (HCl). Rock-Eval pyrolys is was conducted using an OGE-VI Rock

Pyrolysis instrument based on the procedures described by Espitalié et al. (1977). The kerogen

elements were determined using an Elementar vario MARCO CHNS elemental analyzer. The

f
oo
sample powders were pretreated with HCl and a mixed solution of HCl and hydrofluoric acid (HF)

to remove carbonate and silicate minerals. The residue was separated by heavy liquids , and
pr
chloroform (CHCl3 ) was used to extract the soluble components. Whole-rock mineralogical
e-
analyses were measured using a D8 DISCOVER X-Ray Diffractometer (XRD) to provide

semi-quantitative relative abundances of various minerals.


Pr

The vitrinite reflectance (Ro) was analysed using a ×50 oil immersion objective lens and a

Leitz MPV-SP microphotometer, following the procedures described by Schoenherr et al. (2007).
al

A total of 17 samples from wells ST1H, J94, J97, J98X, J116X, and JG13 were selected for
rn

biomarker analyses referring various units of Es 3L . Sample types included marlstones,

hydrocarbons stored in rudstones, and oil samples from production. The gas chromatographic
u

analyses of the saturated hydrocarbon fraction were performed with a HP 6890 chromatograph
Jo

equipped with an HP-5 column (30 m × 0.32 mm i.d.; film thickness 0.25 μm). The oven

temperature was initially set at 35 ℃ for 5 min, increased to 325 ℃ at 3 ℃/min and maintained for

20 min. Helium was utilised as the carrier gas. Gas chromatography-mass spectrometry analyses

of the saturated hydrocarbon fraction were carried out on an Agilent 5975i instrument interfaced to

a HP 6890 chromatograph equipped with the same type of column used during the gas

chromatographic analyses. The oven temperature was initially set at 50 ℃ for 1 min, increased to

120 ℃ at 20 ℃/min, to 310 ℃ at 4℃/min, and maintained for 30 min. Helium was utilised as the

carrier gas again. The abundance ratios of selected biomarkers were calculated from peak areas in

the relevant mass chromatograms.

5
Journal Pre-proof

4. Results
4.1. Lithology characteristics

Five fine-grained rock lithofacies and three rudstone lithofacies have been identified in

previous studies, including varve-like, laminated marlstones, non-regular laminated marlstones,

interlaminated marlstone-siltstones, massive marlstones, massive marlstone-siltstones,

clast-supported carbonate rudstones, matrix-supported carbonate rudstones, and mixed-source

carbonate rudstones (Kong et al., 2017, 2019a). These lithofacies can be grouped into two types

according to their origins: normal sediment facies and event sediment facies (Kong et al., 2017).

Therefore, we simplified the eight lithofacies into three groups: (1) varve-like, laminated

f
oo
marlstones; (2) graded laminated to massive marlstones and siltstones; and (3) carbonate rudstones.

The varve-like, laminated marlstones are characterised by alternating light, calcite laminae and
pr
dark, clay laminae. It represents normal, suspended sediment that was controlled by the seasonal
e-
plankton activity and variable material input, and deposited in a still, anoxic environment (Kong et

al., 2017). The graded laminated to massive marlstones and siltstones are characterised by a very
Pr

faint lamination or a fining upwards lamination and correspond to deposits developed from

turbidity currents (Kong et al., 2017). The carbonate rudstones are derived from the western
al

Ningjin Uplift by mechanical transportation (Jiang et al., 2007). Their distribution characteristics
rn

are shown in Fig. 3.

4.2. Abundances of organic matter


u

The richness of organic matter in wells ST1H and J116X shows significant fluctuation among
Jo

different units (Fig. 3). The highest TOC contents occur in Unit II L, with an average of 2.62 wt%

in Well ST1H, and an average of 2.44 wt% in Well J116X (Fig. 4). Of the remaining units, II U and

VL have higher TOC contents (Fig. 4). Units I L and VU display the lowest TOC contents, and their

average TOC contents are all less than 1.0 wt% (Fig. 4).

The Rock-Eval parameter S2 and TOC can be used to evaluate the hydrocarbon generation

potential of source rocks (Peters and Cassa, 1994). The average values of TOC and S 2 indicate that

most of the samples from Unit II L are very good source rocks, whereas samples from units I L and

VU indicate that they are fair to poor source rocks (Fig. 5). The results show that the remaining

units are good source rocks (Fig. 5). The variation in the source rock quality in wells ST1H and

J116X is basically the same as the variation in the richness of organic matter.
6
Journal Pre-proof

4.3. Maturity and types of organic matter

The Tma x values of Well ST1H range from 424 to 452 ℃ with an average of 444 ℃ (Fig. 6A),

which is in the range of mature zone of hydrocarbon generation (Espitalié et al., 1986). The

vitrinite reflection (Ro) values of Well ST1H range from 0.38 to 0.62, indicating them to be

immature to early mature (Fig. 7). The lower Ro values than expected may be influenced by the

composition of kerogen maceral (Hao and Chen, 1992; Zhao et al., 2014).

The plot of HI versus Tmax indicates that types of organic matter vary with various units in

Well ST1H (Fig. 6A). Most of organic matters of Units IIL , IIU, and VL belong to type I. In

contrast, units IL and VU mainly contain type III organic matter. The remaining units contain

f
oo
organic matter including type II and transitional type between type II and type III. Organic matter

types can also be identified by atomic H/C versus O/C (Van Krevelen, 1961; Tissot et al., 1974).
pr
The results show that kerogen of Well ST1H is mainly Type I to Type II (Fig. 6B). Even though
e-
the results of HI versus Tmax and atomic H/C versus O/C have a difference in the range of organic

matter types, all results show that changes of organic matter types follow a consistent pattern. The
Pr

origins of organic matter inputs of various units are discussed below.

4.4. Biomolecular markers


al

4.4.1. Normal alkanes and isoprenoids


rn

Normal alkanes and isoprenoids are the dominant compounds in sedimentary rocks. C15-C35

n-alkanes are present as shown in Fig. 7. Samples from different units of Es 3 L show unimodal
u

distribution of n-alkanes with the maxima changing from nC17 to nC25 (Fig. 7). The
Jo

terrigenous/aquatic ratio (TAR) can be used to identify changes in the relative amounts of

terrigenous and aquatic organic matters (Bourbonniere and Meyers, 1996). The results show that

samples from units III L, IIIU, IV, and VU have relatively higher TAR values (average value of 2.51)

than the other samples (average value of 0.81) (Table 1). Ratio of nC21-/nC22 + is biomarker

parameter used to analyse oil-source correlation in the study area (Li et al., 2017). Figure 8A

shows that nC21 -/nC22+ ratios have a negative correlation with TAR values such that nC21 -/nC22 +

ratios decrease with the increase in TAR values. This may be because the two parameters are

contrasting in their response to land plant inputs (Peters et al., 2005). Carbon preference index

(CPI) and odd-to-even predominance (OEP) provides information regrading organic matter source,

depositional environment, and thermal maturity (Peters et al., 2005). CPI values indicate that
7
Journal Pre-proof

majority of the samples from units I L , IU, and IIL have an even-carbon preference, apart from

samples J94-1, and J98-1 (Table 1). Sample ST1H-5 has a distinct odd-carbon preference with CPI

and OEP values at 1.35 and 1.59, respectively (Table 1). Other samples have CPI values of

approximately 1.0 (Table 1). Pristane (Pr) to phytane (Ph) ratios vary with samples, with high

values ranging from 1.14 to 1.71 (samples ST1H-4, ST1H-5, ST1H-6, ST1H-8, J94-2, J98-2, and

JG13-1) and low values ranging from 0.33 to 0.53 (samples ST1H-1, ST1H-2, ST1H-3, ST1H-9,

and J97-1) (Fig. 9A). The ratio of isoprenoids to n-alkanes has been calculated, with 0.26–0.88 for

Pr/nC17 and 0.30–1.48 for Ph/nC18 (Table 1). The plot of Pr/nC17 versus Ph/nC18 can be used to

indicate redox conditions during deposition (Shanmugam, 1985).

f
oo
4.4.2. Hopanes

The hopane biomarkers are dominated by the presence of 18α-trisnorhopane (Ts),


pr
17α-trisnorhopane (Tm), C30 -hopane, C29 -norphane, and a certain quantity of homohopanes
e-
(C31 -C35) (Fig. 7). Distribution of homohopanes is dominated by C31 homohopane and decreases

from C31 homohopane to C35 homohopane (Fig. 7). Ts/(Tm+Ts) ratios from samples of Well ST1H
Pr

increase with depth except for Sample ST1H-5 (Table 2). Moreover, C30 M/C30 H ratios from the

same samples show a tendency to decrease with increase in depth (Table 2). The values of two
al

parameters from other samples are shown in Table 2 and are discussed below. Gammacerane is a
rn

peculiarity of the m/z 191 chromatogram. Samples from units IL, IU, and VU of Es 3L have high

Gammacerane/C31 homohopane ratios ranging from 0.35 to 0.94. Gammacerane/C31 homohopane


u

ratios of samples from Unit II L range from 0.20 to 0.29. Other samples have low
Jo

Gammacerane/C31 homohopane ratios ranging from 0.08 to 0.11; the data are presented in Table 2.

4.4.3. Steranes

C27 , C28 , and C29 steranes are the most common sterane homologues of regular steranes and

diasteranes. The relative abundance of C27 , C28, and C29 steranes shows sensitivity to organic

matter type (Huang and Meinschein, 1979). Their abundance in majority of the samples exhibit a

roughly “V” type distribution (Fig. 7). The ternary diagram of C 27, C28 , and C29 steranes from

samples of Well ST1H suggests that marlstones in this study are characterised by organic matter

derived from a mixture of plankton and land plants (Fig. 10). Figure 8B shows that ratios of

C27 /C29 sterane of most samples have negative correlation with TAR values, while some samples

are exceptions such as samples J98-1, J116X-1, and JG13-1. The ratios of C29 20S/(20S+20R) and
8
Journal Pre-proof

C29 ββ/(αα+ββ) are in the range of 0.17–0.50 and 0.28–0.47, respectively (Table 2).

5. Discussion
L
5.1. Evolution of the thermal maturity of Es 3

Biological markers can be used to assess the maturity of source rocks (Gürgey, 1999).
L
Variations in the vertical thermal maturity associated with biomarkers related to Es 3 are present in

the marlstone samples from Well ST1H (Fig. 11). Isomerization in the C29 regular sterane causes

increases in the 20S/(20S+20R) and ββ/(αα+ββ) ratios, as thermal maturity increases at a certain

stage of oil generation (Mackenzie, 1984; Huang et al., 1990; Han et al., 2017). The ratios of the

f
oo
L
two parameters show a similar tendency in which both have low values in upper units of Es 3 (VU ,

VL , and IV) and reach relatively stable high values with increasing burial depth (Fig. 11).
pr
Differences in thermal stability cause the Ts/(Ts+Tm) values to increase, and the C30 M/C30 H
e-
values to decrease with increasing thermal maturity (Seifert and Moldowan, 1978; Peters et al.,

2005). The results suggest that the Ts/(Ts+Tm) values of marlstone samples in Well ST1H are
Pr

L
relatively low in the upper and middle units of Es 3 (from VU to IIU) and high in the lower units of

Es 3 L (IIL, IU , and IL) (Fig. 11). Accordingly, the C30 M/C30 H values are relatively higher in upper
al

units of Es 3 L (VU , VL , and IV) and decrease gradually in the middle units of Es 3 L (IIIU , IIIL , and
rn

IIU ), reaching a low stable value in lower units of Es 3 L (IIL , IU , and IL ) (Fig. 11). Maturity markers

Ro and Tmax also tend to increase as depth increases (Fig. 11). The abnormally low Tmax values of
u

samples from units I L and IU are due to low abundance of S2 , which lead to inaccurate Tmax values.
Jo

L
Combined with various maturity parameters, we propose that Es 3 strata were not influenced by

late tectonic activity because all the parameters have obvious correlations with burial depth. All

the Es 3 L strata may be located in relatively mature zones based on values of Tmax and CPI and OEP

as shown in previous studies (Zhao et al., 2014). Most CPI and OEP values are near 1, suggesting
L
that they have been modified by maturity (Peters et al., 2005). The Es 3 strata of Well ST1H area

can be divided into three groups based on the evolution of their maturity. Group I include units VU ,

VL , and IV, and are characterised by relatively low thermal maturity evolution, as shown in Fig. 11.

Group II include units III U, IIIL, and II U, and represent the transition from low maturity to maturity.

Group III include II L , IU , and IL and have obvious characteristics of relatively high maturity. We

adopted this scheme as the basic premise in the following discussion on sources of organic matter
9
Journal Pre-proof

and maturity heterogeneity caused by hydrocarbon migrations.

5.2. Sedimentary environments and sources of organic matter

Variation of depositional environments of Es3 L can be indicated by biomarkers of Ga/C31 H

and Pr/Ph. The ratio of Ga/C31 H is a salinity index used to show the relative change in salinity of a

water body (Peters et al., 2005). Figure 12 indicates that the hydrological properties of Es 3 L strata

in the basin underwent changes in salinity, from relatively high to low, and the increasing to

relatively high. The Pr/Ph ratio can be used to reflect redox conditions (Powell and McKirdy, 1973;

Didyk et al., 1978; Peters et al., 2005); Pr/Ph values less than 0.8 indicate anoxic conditions

during deposition (Chen et al., 1996), whereas values ranging from 1.0 to 2.0 have been reported

f
oo
in anoxic-subanoxic and fresh-brackish water environments (Chen et al., 1996). Figure 9A

indicates that the Pr/Ph ratios of marlstone samples have a negative correlation with Ga/C 31 H,
pr
suggesting reducing conditions that increase with salinity (Peters and Moldowan, 1993). This
e-
conclusion can be further supported by Fig. 9B, which shows that most of samples with high

salinity are located in the range of the reductive environment. According to the two parameters of
Pr

samples ST1H-2, J97-2, and JG13-1, Unit I U is characterised by relatively high salinity and its

anoxic conditions are gradually noticeable from the basin edge to basin centre (Table 1). The high
al

salinity of the unit may be caused by strong evaporation and relatively low precipitation during the
rn

initial stage (Jiang et al., 2007; Kong et al., 2017). Dramatic changes in the Ga/C31 H ratios

between Es 3L strata and Es 3 U were found in a previous study (Li et al., 2017), which revealed that
u

Ga/C31 H ratios of Es 3U samples are in the range of 1.0–4.0. Considering that no obvious tectonic
Jo

L U
change occurred in the basin during the stage between Es 3 and Es 3 (Jiang et al., 2007) and that

there was an intermittent increase in the dry climate from Es 3 to Es 1 according to palynological

and paleontological data from Shulu Sag (Ren, 1986), we propose that the increase in salinity may

be related to the climate. Therefore, the high value of Ga/C31 H of Unit VU in Well ST1H indicates
L U
that the unit is the transition between Es 3 and Es 3 (Fig. 12). Unit II L has a relatively lower

salinity than Unit I U based on Ga/C31 H ratios (Table 2). However, their Pr/Ph ratios are similar,

indicating similar reducing conditions (Table 1). The lake water level likely increased significantly

in Unit II L , which would have been beneficial to the formation of varve-like, laminated marlstones

(Kong et al., 2017). The stratified water column formed in Unit II L may have been caused by both

temperature and salinity (Sinninghe Damsté et al., 1995). Sample ST1H-3, which has a relatively
10
Journal Pre-proof

higher salinity, has a lower Pr/Ph ratio than Sample ST1H-8 even though they have the same

lithologic association (Fig. 12). Terrestrial inputs can increase Pr/Ph ratios because event

sediments can carry a certain amount of oxygen to the lake bottom (Limbach, 1975; Liang et al.,

2016). The injection of fresh water can also reduce salinity of the waterbody. Therefore, other

units formed in an oxidising or transitional environment have low Ga/C31 H values and high Pr/Ph

ratios (Fig. 9B, 12).

Previous studies suggested that the organic matter in marlstones from Es 3L in Shulu Sag were

derived from both plankton and land plants (Li et al., 2017; Tang et al., 2018). Influences from

organic matter sources on hydrocarbon accumulation were ignored because organic matters in

f
oo
L
Es 3 have similar characteristics (Tang et al., 2018). Even though the relative abundances of C 27,

C28 , and C29 steranes supporting the organic matter in Well ST1H are mixed-source type (Fig. 10),
pr
differences in organic matter compositions can also be presented by pyrolysis parameters (HI and
e-
Tma x), kerogen elements, and C27/C29 steranes ratios (Figs. 6, 8B). The n-alkane distribution of

saturated hydrocarbons can be used to deduce the sources of organic matter (Brassell et al., 1978;
Pr

Hakimi et al., 2016). Higher TAR values indicate more input from land materials, compared to

aquatic sources, from the surrounding watershed (Bourbonniere and Meyers, 1996; Lai et al.,
al

2018). The plot of TAR versus nC21 -/nC22 + can also be used to evaluate the differences in
rn

abundances of light and heavy hydrocarbons among various samples (Fig. 8A). However, thermal

cracking may change the relative abundance of molecular homologs to influence TAR and
u

nC21 -/nC22 + ratios (Peters et al., 2005). Biomarkers of samples from units IU , IIL , and IIU of Well
Jo

ST1H may have been affected by thermal maturity because their TAR ratios are low, and they

have similar values (Fig. 12). Samples ST1H-3 and ST1H-8 have similar lithological associations;

however, the TAR ratio of the former is lower than the latter, possibly in relation to the thermal

cracking of long chain n-alkane. We chose a low maturity sample (J94-3) from Unit II L for the

comparison, which indicated that values of TAR and nC21 -/nC22 + between J94-3 and ST1H-3 were

similar (Table 2). The results suggest that even though thermal cracking may modify values of

biomarkers (e.g. TAR ratio of ST1H-4), the organic matter from units I U and IIL still show a

predominantly aquatic source. This can be deduced from the low maturity samples from same

units and the ratios of the C27 /C29 steranes. High C27 /C29 ratios can reflect a predominantly

plankton/algal source (Huang and Meinschein, 1979; Hakimi et al., 2016). Therefore, values of
11
Journal Pre-proof

C27 /C29 and TAR of most samples show the inverse relationship (Fig. 8B). Samples from units IU

and IIL have C30 sterane, possibly in the form of 4-methyl steranes, which suggests a contribution

from algae (e.g., dinoflagellates) to the sediments (Goodwin et al., 1988). This finding further

supports the idea that units I U and IIL had aquatic sourced organic matter. Abnormally low CPI

values for these units may also indicate the presence of organic matter derived from aquatic

organisms under hypersaline environments, given their maturity (Palacas et al., 1984; Peters et al.,

2005). Figure 6 shows that organic matter within Unit I U may have come from mixed sources,

suggesting that the unit had a certain amount of higher plant input. Sample JG13-1 from this unit

may have been affected by terrigenous organic matter inputs and thus had a relatively low C27 /C29

f
oo
ratio and a relatively high Pr/Ph ratio (Fig. 9A); this is because Well JG13 is near the eastern steep

slope area where nearshore subaqueous fan systems developed (Jiang et al., 2007). The sample
pr
may also have been influenced by thermal cracking, which would have caused the low TAR ratio,
e-
given its maturity (Table 2). We speculated that some hydrocarbons within Unit I U may have

migrated from Unit II L because the low TOC content of Unit I U would have limited its
Pr

hydrocarbon generation potential; this would account for the sample from Unit I U (ST1H-2)

having characteristics similar to those seen in the gas chromatograms and mass chromatograms in
al

a sample from Unit II L (ST1H-3) (Fig. 7). Therefore, it is not possible to determine the organic
rn

matter source of Unit I U based on available data. Thus, considering multiple parameters, we

propose that Unit II L has more aquatic organic matters than other units.
u
Jo

5.3. Accumulation of organic matter

The accumulation of organic matter in marlstones in Es 3 L is controlled by both organic matter

sources and depositional environments. The deposits that filled Es 3 L in the Shulu Sag were mainly

source-controlled, in that abundant terrigenous clasts were transported into the basin from

catchment given the tectonic activity occurring at the time (Jiang et al., 2007; Zheng et al., 2015).

Previous lithologic analyses have suggested that there were more event sediments present in the

basin at the time compared to the normal background sediments (Jiang et al., 2007; Kong et al.,

2017). Variations in vertical of representation of organic matter sources are seen in the ratios of

C27 /C29 sterane in Fig. 12 and in the other parameters discussed in section 5.2. When the terrestrial

input increased, the TOC contents decreased (e.g. units III U and IV), which may have been related
12
Journal Pre-proof

to the dilution of inorganic material and adverse preservation conditions (Liang et al., 2016; Wu et

al., 2019). Salinity has a complex relationship with the accumulation of organic matter (Hammer,

1981; Li et al., 2015; Yang and Schulz, 2019). Increases in salinity do not necessarily lead to

decreases in primary productivity, because the number of individuals surviving within a particular

species can increase even though overall species diversity decreases (Clark and Philp, 1989).

Comparative study conducted by Horsfield et al. (1994) indicates that source rocks deposited in an

alkaline lake have better potential than those in a freshwater lake. Studies also suggest that the

salinity stratification of water column is beneficial for preserving organic matter (Kirkland and

Evans, 1981; Clark and Philp, 1989). However, the degradation caused by sulfate-reducing

f
oo
bacteria in a saline lake can result in a loss of organic matter (Didyk et al., 1978; Kelts, 1988; Katz,

2001). When the waterbody had a high salinity in units I L , IU, and VU , the organic matter content
pr
in the sediments were low (Fig. 4, Fig. 12). However, the highest organic matter content in the
e-
study area was found in Unit II L under conditions of moderate salinity, rather than in other units

under low salinity conditions (Fig. 12). Therefore, organic matter accumulation is likely the result
Pr

of interaction of productivity and preservation (Tyson, 2005). The details of how organic matter

accumulated in various units in Es 3 L of the Shulu Sag are discussed below.


al

During the initial rift stage (Unit I L ), the lake basin was narrow and shallow (Kong et al.,
rn

2017). Abundant rudstones that formed after sudden floods or tectonic events occurred in the

bottom of the basin (Liu et al., 2017). Following the rudstone deposits, the basin developed
u

greyish-green shallow water facies marlstone-siltstones (Kong et al., 2017). The primary
Jo

productivity of the unit was at its lowest, based on its TOC content and the type of organic matter

present (Figs. 4, 6). This suggests that a high input of terrigenous materials and oxic environments

did not benefit the accumulation of organic matter in the Shulu Sag.

Unit IU mainly developed grey coloured, non-regular laminated marlstones, suggesting that

lake water level obviously increased there compared to in Unit I L . The number of aquatic

organisms may have increased in a high salinity environment; however, the relatively low TOC

content of the unit may be related to the degradation of organic matter caused by bacteria (Katz,

2001). Therefore, the influence of primary productivity on the accumulation of organic matter is

limited when a water body was highly saline.

For Unit II L, the lake water level rose further, and salinity decreased accordingly. The unit
13
Journal Pre-proof

developed varve-like, laminated marlstones with alternating calcite laminae and clay laminae

(Kong et al., 2017). The rhythm feature supported algal blooms such that the formation of calcite

laminae was induced by biochemical processes (Kong et al., 2017). The suitable salinity did not

damage preservation conditions of organic matter and maintained the reductive environment. Such

environments were also beneficial for the preservation of varve textures (Anderson, 1986).

Therefore, a high organic matter content is the result of high productivity and a strong reductive

condition caused by suitable salinity or temperature.

From Unit II U to IV, the basin was in an active stage of tectonic activity and seismic

development was abundant (Zheng et al., 2015; Kong et al., 2019b). The rudstones interval in Unit

f
oo
IIIL is considered to be related to paleo-earthquakes (Liu et al., 2017). Therefore, few varve-like,

laminated marlstones occur in these units (Fig. 3). Low Ga/C 31 H values suggest that these units
pr
had a low salinity (Fig. 12). Figure 12 shows that TOC values varied with the values of TAR,
e-
C27 /C29 sterane, and Pr/Ph. The results suggest that enrichment by organic matter is controlled by

terrigenous material input, primary productivity, and redox conditions. The abundant terrigenous
Pr

materials may have been brought into the basin by bottom flows, which could also have carried a

large amount of oxygen that would have affected reductive conditions (Liang et al., 2016).
al

Terrestrial inorganic material inputs could have diluted the accumulation of organic matter and led
rn

to the low TOC values observed today. Therefore, an increased input of terrestrial materials did

not benefit the accumulation of organic matter in the basin.


u

The Unit VL was a stage in which the water level of the lake rose again, and the basin was a
Jo

still environment with weak inputs of terrestrial material. The environment caused the unit to

develop varve-like, laminated marlstones. However, the low salinity caused the unit to have worse

preservation conditions than those in Unit II L . Planktonic organic matter may have been oxidised

in the upper water column because TOC values of this unit were relatively low.

Unit VU had a similar depositional environment as Unit I U , according to the Ga/C31 H

biomarker (Fig. 12). However, rapid salinization may have caused a rapid die-off in aquatic

organisms as they would not have been able to adapt to the changes in their environment. The

sudden change may have been related to climatic conditions, as discussed in section 5.2. The

decrease in aquatic organisms affected the percentage of terrestrial organic matter; therefore, the

TAR value is high and the C27 /C29 sterane ratio is low. Thus, this unit presents low TOC values
14
Journal Pre-proof

due to the low primary productivity level and the unfavourable preservation conditions caused by

the sudden increase in salinity.

5.4. Hydrocarbon accumulation models

Previous studies have suggested that source rocks from Es3 L in the Shulu Sag were all in the

mature zone (Zhao et al., 2014), which is supported by Tmax values of various units (Fig. 6A).

Therefore, all the source rocks from Es3 L have hydrocarbon generation potentials to a certain

extent. Thermal maturity profile of marlstones along the depth (Fig. 11) suggests that

hydrocarbons within marlstones belong to retention resources. In situ accumulation is considered

f
oo
to be the primary model of hydrocarbon accumulation in previous studies (Li et al., 2017), because

both marlstones and rudstones in Es 3 L have very low porosity (0.4–2.6 %) and permeability (0.04–
pr
17.1 md) (Li et al., 2017; Kong et al., 2019a). Short distance hydrocarbon migration from source
e-
rocks to rudstones was responsible for formation of oil production sweet pots (Tang et al., 2018).

According to the short distance migration model, hydrocarbons from rudstones should have
Pr

similar biomarker characteristics as nearby marlstones. Otherwise, a significant difference in

biomarkers between rudstones and their nearby source rocks should be found. Values of
al

Ts/(Ts+Tm), CPI, and OEP from Sample ST1H-5 are obviously higher than those of nearby
rn

marlstones (Fig. 12). The ratio of Ts/(Ts+Tm) depends on maturity, as well as source and

depositional environment (Moldowan et al., 1986). A possible reason for abnormal values of the
u

ratio from Sample ST1H-5 is that hydrocarbons in Unit III L rudstones of Well ST1H were derived
Jo

from more mature organic matter. The influence of the lithology of the source rocks can be

ignored as all the source rocks in Es 3 L are rich in carbonate. Salinity may influence the ratio of

Ts/(Ts+Tm) (Fan et al., 1987) as Sample ST1H-9 has higher value of the biomarker than Sample

ST1H-8 (Table 2). However, thermal maturity was the main factor influenc ing the biomarker with

respect to the values of Ts/(Ts+Tm) of units I U and VU . Abnormal high values of CPI and OEP

from Sample ST1H-5 suggest that organic matter is derived mainly from land plants (Bary and an

Evans, 1961), based on its Ro value, corresponding to its depositional background of strong

terrigenous material inputs. Therefore, vertically upward migration of hydrocarbons in Es 3 L of

Shulu Sag can be speculated.

Figure 13 shows the relationship of maturity biomarkers between source rocks and rudstones
15
Journal Pre-proof

and oils from production. As stated above, hydrocarbons from Unit III L (Sample ST1H-5) belong

to mature source rocks (Fig. 13). Hydrocarbons from rudstone sample ST1H-1 (Unit I L ) have

similar maturity biomarkers as mature source rocks (Fig. 13), suggesting they are migrated from

nearby source rocks with abundant organic matter content. However, hydrocarbons from rudstone

samples of Well J94 have greater differences in maturity biomarkers. Hydrocarbons of Sample

J94-1 (Unit I L ) may be migrated from mature source rocks in deeper strata, while hydrocarbons of

Sample J94-2 (Unit I U ) may from low mature source rocks in its overlying unit because Sample

J94-2 has similar maturity biomarkers as Sample J94-3 (Unit II L ) (Fig. 13). Moreover, the

differences in maturity biomarkers between rudstone sample J97-1 (Unit I L ) and marlstone sample

f
oo
(Unit I U ) suggest that hydrocarbons from upper low maturity source rocks were also migrated into

lower rudstones reservoirs (Fig. 13). Two oil samples achieved by production show high maturity
pr
and have similar characteristic gas chromatograms and mass chromatograms as compared to

mature source rocks (Fig. 7) suggesting that main hydrocarbons in Es 3 L were migrated from
e-
mature source rocks.
Pr

L
A new hydrocarbon accumulation model of Es 3 submember in the Shulu Sag is proposed

(Fig. 14). Hydrocarbons in the study area can be divided into two types depending on burial depth:
al

mature and low mature. Based on the organic matter sources analyses thus far, varve-like,
rn

laminated marlstones have highest TOC contents; therefore, these marlstones located in the deep

strata have the best hydrocarbon generation quality and potential for the formation of in situ
u

marlstone oil reservoirs. Hydrocarbons can migrate into adjacent rudstones to form rudstone oil
Jo

reservoirs; these reservoirs have similar maturity characteristics as their neighbouring source rocks.

Numerical simulation of the stress field suggests that the lower gentle slope developed abundant

structural fractures (Wang et al., 2007). The slope break linking the gentle slope and subsag is

located in the release zone of the tensile stress, which was conducive to formation of fractures. On

the other hand, the Shulu Sag developed multiple NE-SW trending faults in the gentle slope,

which may have caused secondary faults during tectonically active s tages. These fractures

provided effective channels for long distance migration of hydrocarbons. Solid residual bitumen

can be found in fractures as evidence of hydrocarbon migration (Fig. 15). Therefore,

unconventional hydrocarbon systems in Es 3L of the Shulu Sag have multiple oil accumulation

ways: (1) retained in situ; (2) migration over short distances; and (3) migration over long distances.
16
Journal Pre-proof

L
Due to the lower maturity and organic matter content, the upper strata of Es 3 have lower

hydrocarbon generation potential than the lower strata. Therefore, previous studies have shown

that the lower strata have better production potential than the upper strata (Tang et al., 2018; Kong

et al., 2019a). Even though some low mature hydrocarbons could migrate into lower rudstone

deposits through fractures, they may play a limited role in effective oil reservoirs, as evidenced by

the oil samples obtained. In contrast, hydrocarbons released from mature source rocks within

lower strata could migrate through continuous rudstone deposits with multiple fractures from

deeper strata to the gentle slope, which could lead to potential hydrocarbon accumulation. Organic

matter sources, preservation conditions, maturity, and hydrocarbon migration ways jointly

f
oo
L
controlled the formation of oil reservoirs in Es 3 of the Shulu Sag. Therefore, deeper strata with

high TOC contents and effective migration channels is considered to be the target area with

highest exploration potential. pr


e-
6. Conclusions
Pr

L
The Es 3 submember in the Shulu Sag has three primary lithologic assemblages, including: (1)

varve-like, laminated marlstones; (2) graded laminated to massive marlstones and siltstones; and
al

(3) carbonate rudstones. The submember can be subdivided into nine units: I L , IU , IIL , IIU , IIIL , IIIU,
rn

IV, VL , and VU . Varve-like, laminated marlstones are mainly developed in units II L and VL .

Carbonate rudstones are mainly developed in units I L and IIIL .


u

Organic matter types vary with the units of Es3 L in the Shulu Sag. Normal alkane
Jo

distributions combined with other biomarkers suggest that compared to other units, Unit II L

possesses organic matter that is primarily aquatic in origin. The depositional filling of Es 3 L in the

Shulu Sag was controlled by basin evolution. During the depositional period of Unit II L , the basin

activity was at a stable stage with suitable salinity, which developed abundant varve-like,

laminated marlstones with low terrestrial clasts input. High primary productivity and good

preservation conditions were beneficial for organic matter accumulation. Therefore, source rocks

of Unit II L have the best hydrocarbon generation potential.

Analyses of biomarkers from marlstone samples suggest that the thermal maturity of the

source rocks increases with burial depth. The heterogeneity of thermal maturity of hydrocarbons

within the rudstone samples suggests that different hydrocarbon migration patterns existed within
17
Journal Pre-proof

L
the Es 3 submember. In situ marlstone oil reservoirs were developed in organic matter-rich

marlstone intervals in which hydrocarbons were retained in matrix pores due to low porosity and

permeability. Short distance hydrocarbon migration occurred in rudstone intervals interbedded

with organic matter-rich marlstones. Hydrocarbons can also migrate over a long distance through

fractures. Potential high-quality oil reservoirs are formed in deeper strata having mature, high

TOC source rocks and effective rudstone reservoirs.

Acknowledgements

f
This work was supported by the National Major Research Program for Science and

oo
Technology of China [Grant No. 2017ZX05009-002], and the Natural Science Foundation [Grant

No. 41772090]. We thank editor-in-chief and two anonymous reviewwes for their insightful
pr
comments, which have significantly improved our initial manuscript.
e-
References
Pr

[1] Anderson, R.Y., 1986. The Varve Microcosm: Propagator of Cyclic Bedding.

Paleoceanography 1, 373–382.
al

[2] Bary, E.E., an Evans, E.D., 1961. Distribution of n-paraffins as a clue to the recognition of
rn

source beds. Geochim. Cosmochim. Acta 22, 2–9.


u

[3] Bourbonniere, R.A., Meyers, P.A., 1996. Sedimentary geolipid records of historical changes
Jo

in the watersheds and productivities of Lakes Ontario and Erie. Limnol. Oceanogr. 41, 352–

359.

[4] 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),

Aquatic pollutants: transformation and biological effects. Oxford: Pergamon, pp. 69-86.

[5] Chen, J., Bi, Y., Zhang, J., 1996. Oil-source correlation in the Fulin basin, Shengli petroleum

province, East China. Org. Geochem. 24, 931–940.

[6] Clark, J.P., Philp, R.P., 1989. Geochemical Characterization of Evaporite and Carbonate

Depositional Environments and Correlation of Associated Crude Oils in the Black Creek

Basin, Alberta. Bull. Can. Petrol. Geol. 37(4), 401-416.

18
Journal Pre-proof

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

indicators of palaeoenvironmental conditions of sedimentation. Nature 272, 216–222.

[8] Espitalié, J., Deroo, G., Marquis, F., 1986. Rock-Eval pyrolysis and its applications; 3. Rev. l.

Fr. Petrol. 41, 73–89.

[9] Espitalié, J., Madec, M., Tissot, B., Mennig, J., Leplat, P., 1977. Source rock characterization

method for petroleum exploration. Offshore Technology Conference, Houston, Texas.

[10] Fan, P., King, J. D. and Claypool, G. E., 1987. Characteristics of biomarker compounds in

Chinese crude oils. In: Kumar, R. K., Dwivedi, P., Banerjie, V., Gupta, V. (eds), Petroleum

Geochemistry and Exploration in the Afro-Asian Region, Balkema, Rotterdam, pp. 197-202.

f
oo
[11] Goodwin, N.S., Mann., LS., Patience, R.L., 1988. Structure and significance of 4-methyl

steranes in lacustrine shales and oils. Org. Geochem. 12, 495–506.


pr
[12] Guo, Y., Song, Y., Fang, X., Pang, X., Li, T., 2018. Reservoir characterization of an
e-
organic-rich dolomitic tight-oil reservoir, the Lower Cretaceous Xiagou Formation in the

Qingxi Sag, Jiuquan Basin, NW China. Mar. Petrol. Geol. 89, 541–559.
Pr

[13] Gürgey, K., 1999. Geochemical Characteristics and Thermal Maturity of Oils from the

Thrace Basin (Western Turkey) and Western Turkmenistan. J. Petrol. Geol. 22, 167–189.
al

[14] Hakimi, M.H., Abdullah, W.H., Alqudah, M., Makeen, Y.M., Mustapha, K.A., 2016. Organic
rn

geochemical and petrographic characteristics of the oil shales in the Lajjun area, Central

Jordan: Origin of organic matter input and preservation conditions. Fuel 181, 34–45.
u

[15] Hammer, U. T., 1981. Primary production in saline lakes. A review. Hydrobiologia, 81: 47–
Jo

57.

[16] Han, C., Jiang, Z., Han, M., Wu, M., Lin, W., 2016. The lithofacies and reservoir

characteristics of the Upper Ordovician and Lower Silurian black shale in the Southern

Sichuan Basin and its periphery, China. Mar. Petrol. Geol. 75, 181–191.

[17] Han, Y., Horsfield, B., Curry, D.J., 2017. Control of facies, maturation and primary migration

on biomarkers in the Barnett Shale sequence in the Marathon 1 Mesquite well, Texas. Mar.

Petrol. Geol. 85, 106–116.

[18] Hao, F., Chen, J.Y., 1992. The cause and mechanism of vitrinite reflectance anomalies. J.

Petroleum Geol. 15, 419–434.

[19] Horsfield, B., Curry, D., Bohacs, K., Littke, R., Rullkötter, J., Schenk, H., Radke, M.,
19
Journal Pre-proof

Schaefer, R., Carroll, A. and Isaksen, G., 1994. Organic geochemistry of freshwater and

alkaline lacustrine sediments in the Green River Formation of the Washakie Basin, Wyoming,

USA. Org. Geochem. 22, 415–440.

[20] Hu, T., Pang, X., Jiang, S., Wang, Q., Zheng, X., Ding, X., Zhao, Y., Zhu, C., Li, H., 2018.

Oil content evaluation of lacustrine organic-rich shalewith strong heterogeneit.: A case study

of the Middle Permian Lucaogou Formation in Jimusaer Sag, Junggar Basin, NW China. Fuel

221, 196–20.

[21] Huang, D., Li, J., Zhang, D., 1990. Maturation sequence of continental crude oils in

hydrocarbon basins in China and its significance. In: Durand, B., Behar, F. (Eds), Org.

f
oo
Geochem. 16, 521–529.

[22] Huang, W.Y., Meinschein, W.G., 1979. Sterols as ecological indicators. Geochim.

Cosmochim. Acta 43, 739–745. pr


e-
[23] Jarvie, D.M., 2012a. Shale resource systems for oil and gas: Part 1—Shale-gas resource

systems. In: Breyer, J.A. (Ed), Shale reservoirs-giant resources for the 21st century, AAPG
Pr

Mem. 97, pp. 69–87.

[24] Jarvie, D.M., 2012b. Shale resource systems for oil and gas: Part 2—Shale-oil resource
al

systems. In: Breyer, J.A. (Ed), Shale reservoirs-giant resources for the 21st century, AAPG
rn

Mem. 97, pp. 89–119.

[25] Jiang, Z., Chen, D., Qiu, L., Liang, H., Ma, J., 2007. Source-controlled carbonates in a small
u

Eocene half-graben lake bas in (Shulu Sag) in central Hebei Province, North China.
Jo

Sedimentology 54, 265–292.

[26] Katz, B.J., 2001. Lacustrine basin hydrocarbon exploration—current thoughts. J. Paleolimnol.

26, 161–179.

[27] 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. Geol. Soc. Special Publication 40: 3–26.

[28] Kirkland, D.W., Evans, R., 1981. Source-Rock Potential of Evaporitic Environment. AAPG

Bull. 65(2), 181–190.

[29] Ko, L.T., Loucks, R.G., Ruppel, S.C., Zhang, T., Peng, S., 2017. Origin and characterization

of Eagle Ford pore networks in the south Texas Upper Cretaceous shelf. AAPG Bull. 101,
20
Journal Pre-proof

387–418.

[30] Kong, X., Jiang, Z., Han, C., Zheng, L., Zhang, J., 2019a. The tight oil of lacustrine

carbonate-rich rocks in the Eocene Shulu Sag: Implications for lithofacies and res ervoir

characteristics. J. Petrol. Sci. Eng. 175, 547–559.

[31] Kong, X., Jiang, Z., Han, C., Li, H., Li, Q., Zheng, L., Yang, Y., Zhang., Xiao, F., 2019b.

Sedimentary characteristics and depositional models of two types of homogenites in an

Eocene continental lake basin, Shulu Sag, eastern China. J. Asian Earth Sci. 179, 165–188.

[32] Kong, X., Jiang, Z., Han, C., Zheng, L., Zhang, Y., Zhang, R., Tian, J., 2017. Genesis and

implications of the composition and sedimentary structure of fine-grained carbonate rocks in

f
oo
the Shulu sag. J. Earth. Sci. 28, 1047–1063.

[33] Lai, H., Li, M., Liu, J., Mao, F., Xiao, H., He, W., Yang, L., 2018. Organic geochemical
pr
characteristics and depositional models of Upper Cretaceous marine source rocks in the
e-
Termit Basin, Niger. Palaeogeogr. Palaeoclimatol. Palaeoecol. 495, 292–308.

[34] Lei, Y., Luo, X., Wang, X., Zhang, L., Jiang, C., Yang, W., Yu, Y., Cheng, M., Zhang, L.,
Pr

2015. Characteristics of silty laminae in Zhangjiatan Shale of southeastern Ordos Basin,

China: implications for shale gas formation. AAPG Bull. 99, 661–687.
al

[35] Li, J., Zhou, S., Li, Y., Ma, Y., Yang, Y., Li, C., 2016. Effect of organic matter on pore
rn

structure of mature lacustrine organic-rich shale: A case study of the Triassic Yanchang shale,

Ordos Basin, China. Fuel 185, 421–431.


u

[36] Li, M., Chen, Z., Cao, T., Ma, X., Liu, X., Li, Z., Jiang, Q., Wu, S., 2018. Expelled oils and
Jo

their impacts on Rock-Eval data interpretation, Eocene Qianjiang Formation in Jianghan

Basin, China. Int. J. Coal Geol. 191, 37–48.

[37] Li, Q., You, X., Jiang, Z., Zhao, X., Zhang, R., 2017. A type of continuous petroleum

accumulation system in the Shulu sag, Bohai Bay bas in, eastern China. AAPG Bull. 101,

1791–1811.

[38] Li, Y., Fan, T., Zhang, J., Wei, X., Zhang, J., 2015. Impact of paleoenvironment, organic

paleoproductivity, and clastic dilution on the formation of organic -rich shales a case study

about the Ordovician-Silurian black shales, southeastern Chongqing, South China. Arabian J.

Geosci. 8, 10225–10239.

[39] Liang, C., Cao, Y., Jiang, Z., Wu, J., Song, G., Wang, Y., 2017. Shale oil potential of
21
Journal Pre-proof

lacustrine black shale in the Eocene Dongying depression: Implications for geochemistry and

reservoir characteristics. AAPG Bull. 101, 1835–1858.

[40] Liang C, Cao, Y., Liu, K., Jiang, Z., Wu, J., Hao, F., 2018a. Diagenetic variation at the lamina

scale in lacustrine organic-rich shales: Implications for hydrocarbon migration and

accumulation. Geochim. Cosmochim. Acta 229, 112–128.

[41] Liang, C., Jiang, Z., Cao, Y., Wu, J., Wang, Y., Hao, F., 2018b. Sedimentary characteristics

and origin of lacustrine organic-rich shales in the salinized Eocene Dongying Depression.

GSA Bull. 130, 154–174.

[42] Liang, C., Jiang, Z., Cao, Y., Wu, M., Guo, L., Zhang, C., 2016. Deep-water depositional

f
oo
mechanisms and significance for unconventional hydrocarbon exploration: A case study from

the lower Silurian Longmaxi shale in the southeastern Sichuan Basin. AAPG Bull. 100, 773–

794. pr
e-
[43] Limbach, G.M., 1975. On the origin of petroleum. Proceeding of Ninth World Petroleum:

Congress Applied Science Publishers, pp. 357–369.


Pr

[44] Liu, X., Zheng, L., Jiang, Z., Kong, X., 2017. Formation mechanisms of rudstones and their

effects on reservoir quality in the Shulu sag, Bohai Bay Basin, Eastern China. J. Earth Sci. 28,
al

1097–1108.
rn

[45] Mackenzie, A.S., 1984. Application of biological markers in the petroleum geochemistry. In:

Brookes, J., Welte, D.H. (Eds), Advances in Organic Geochemistry, London: Academic Press,
u

pp. 115–214.
Jo

[46] Marzi, R., Torkelson, B.E., Olson, R.K., 1993. A revised carbon preference index. Org.

Geochem. 20, 1303–1306.

[47] Mitchell J., 2011. Horizontal drilling of deep granite wash reservoirs, Anadarko Basin,

Oklahoma and Texas. Shale Shak 62, 118–167.

[48] Moldowan, J.M., Fago, F.J., 1986. Structure and significance of a novel rearranged

monoaromatic steroid hydrocarbon in petroleum. Geochim. Cosmochimica. Acta 50, 343–

351.

[49] Mukhopadhyay, P.K., Wade, J.A., Kruge, M.A., 1995. Organic facies and maturation of

Jurassic/Cretaceous rocks, and possible oil-source rock correlation based on pyrolysis of

asphaltenes, Scotian Basin, Canada. Org. Geochem. 22, 85–104.


22
Journal Pre-proof

[50] Palacas, G.P., Anders, D.E., King, J.D., 1984. South Florida Bas in–A prime-example of

carbonate source rocks of petroleum. In: Palacas, J.G. (Ed), Petroleum Geochemistry and

Source Rock Potential of Carbonate Rocks. AAPG Studies in Geology 18, pp. 71–96.

[51] Peters, K.E., Cassa, M.R., 1994. Applied source rock geochemistry: chapter 5: Part II.

Essential elements. In: Magoon, L. B., Dow, W.G. (Eds), The Petroleum System-from Source

to Trap, AAPG Bull. 60, pp. 93–120.

[52] Peters, K.E., Moldowan, J.M., 1993. The Biomarkers Guide: Interpreting molecular fossils in

petroleum fossils and ancient sediments. Englewood Cliffs, New Jersey: Prentice Hall, pp.

363.

f
oo
[53] Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide: Biomarkers and

Isotopes in Petroleum Exploration and Earth History, second ed. Cambridge: vol 2

Cambridge University Press. pr


e-
[54] Powell, T.G., McKirdy, D.M., 1973. Relationship between ratio of pristane to phytane, crude

oil composition and geological environment in Australia. Nat. Phy. Sci. 243, 37–9.
Pr

[55] Ren, Y., 1986. Depositional environments of Shulu Depression-Viewed from the point of

micropaleobotanic florae. Acta Sedimentol. Sinica 4, 101–108 (in Chinese with English
al

Abstract).
rn

[56] Schoenherr, J., Littke, R., Urai, J.L., Kukla, P.A., Rawahi, Z., 2007. Polyphase thermal

evolution in the Infra–Cambrian Ara Group (South Oman Salt Basin) as deduced by solid
u

bitumen maturity. Org Geochem. 38, 1293–1318.


Jo

[57] Seifert, W.K., Moldowan, M.J., 1978. Applications of steranes, terpanes and monoaromatics

to the maturation, migration and source of crude oils. Geochim. Cosmochimica. Acta 42, 77–

95.

[58] Shanmugam, G., 1985. Significance of coniferous rain forests and related organic matter in

generating commercial quantities of oil, Gipps land Bas in, Australia. AAPG Bull. 69, 1241–

1254.

[59] Sinninghe Damsté, J.S., Kenig, F., Koopmans, M.P., Köster, J., Schouten, S., Hayes, J.M., de

Leeuw, J.W., 1995. Evidence for gammacerane as an indicator of water-column stratification.

Geochim. Cosmochim. Acta, 59, 1895–1900.

[60] Smith, M.G., Bustin, R.M., 2000. Late Devonian and Early Mississippian Bakken and
23
Journal Pre-proof

Exshaw black shale source rocks, western Canada sedimentary basin: A sequence

stratigraphic interpretation. AAPG Bull. 84, 940–960.

[61] Tang, X., Zhang, J., Jiang, Z., Zhang, R., Lan, C., Zhao, W., Zhu, J., Wang, J., Zhao, P., 2018.

Heterogeneity of organic-rich lacustrine marlstone succession and their controls to petroleum

expulsion, retention, and migration: A case study in the Shulu Sag, Bohai Bay Basin, China.

Mar. Petrol. Geol. 96, 166–178.

[62] Tissot, B., Durand, B., Espitalie, J., Combaz, A., 1974. Influence ofnature and diagenesis of

organic matter in formation of petroleum. AAPG Bull. 58, 499–506.

[63] Tyson, R.V., 2005. Productivity versus preservation controversy: cause, flaws, and resolution.

f
oo
In: Harris, N.B. (Ed.), The Deposition of Organic Carbon-Rich Sediments: Models,

Mechanisms and Consequences. SEPM (Society for Sedimentary Geology) Special

Publication 82, Tulsa, Oklahoma, pp. 17–34. pr


e-
[64] Van Krevelen, D., 1961. Coal: Typology. Chemistry, Physics, Constitution. Elsevier

Publishing Company, New York.


Pr

[65] Wang, G., Chang, X., Yin, W., Li, Y., Song, T., 2017. Impact of diagenesis on reservoir

quality and heterogeneity of the Upper Triassic Chang 8 tight oil sandstones in the Zhenjing
al

area, Ordos Basin, China. Mar. Petrol. Geol. 83, 84–96.


rn

[66] Wang, M., Cui, Y., Zhang, R., Lu, Y., Wang, S., Lu, L., 2007. Prediction method for marl

fractured reservoir —An example from Shulu Sag. Lithologic Reservoirs 19, 114–119 (in
u

Chinese with English abstract).


Jo

[67] Wu, J., Liang, C., Hu, Z., Yang, R., Xie, J., Wang, R., Zhao, J., 2019. Sedimentation

mechanisms and enrichment of organic matter in the Ordovician Wufeng Formation-Silurian

Longmaxi Formation in the Sichuan Basin. Mar. Petrol. Geol. 101, 556–565.

[68] Xu Z., Liu L., Wang T., Wu K., Dou W., Song X., Feng C., Li X., Ji H., Yang Y., Liu X., 2017.

Characteristics and controlling factors of lacustrine tight oil reservoirs of the Triassic

Yanchang Formation Chang 7 in the Ordos Basin, China. Mar. Petrol. Geol. 82, 265–296.

[69] Yang, S., Schulz, H.M., 2019. Factors controlling the petroleum generation characteristics of

Palaeogene source rocks in the Austrian Molasse Bas in as revealed by principal component

analysis biplots. Mar. Petrol. Geol. 99, 323–336.

[70] Ye, H., Zhang, B., Mao, F., 1987. The Cenozoic tectonic evolution of the Great North China:
24
Journal Pre-proof

two types of rifting and crustal necking in the Great North China and their tectonic

implications. Tectonophysics 133, 217–227.

[71] Zhao, X., Li, Q., Jiang, Z., Zhang, R., Li, H., 2014. Organic geochemistry and reservoir

characterization of the organic matter-rich calcilutite in the Shulu Sag, Bohai Bay Basin,

North China. Mar. Petrol. Geol. 51, 239–255.

[72] Zheng, L., Jiang, Z., Liu, H., Kong, X., Li, H., Jiang, X., 2015. Core evidence of

paleoseismic events in Paleogene deposits of the Shulu Sag in the Boha i Bay Basin, east

China, and their petroleum geologic significance. Sed. Geol. 328, 33–54.

[73] Zou, C., Zhu, R., Chen, Z., Ogg, J.G., Wu, S., Dong, D., Qiu, Z., Wang, Y., Wang, L., Lin, S.,

f
oo
Cui, J., Su, L., Yang, Z., 2019. Organic-matter-rich shales of China. Earth-Sci. Rev. 189, 51–

78.
pr
e-
Pr
al
u rn
Jo

25
Journal Pre-proof

Table 1. Biomarker parameters of n-alkanes from the Shulu Sag.

n-Alkane and isoprenoid ratios


Sample ID Unit Sample
Pr/Ph Pr/nC17 Ph/nC18 TAR nC21 -/nC22 + CPI OEP

ST1H-1
ST1H-2
IL
IU
R
M
0.41
0.51
0.27
0.38
0.51
0.65
1.11
0.98
o f0.63
0.70
0.77
0.84
0.93
0.96
ST1H-3 IIL M 0.50 0.26 0.31

r o
1.01 0.70 0.88 1.02
ST1H-4
ST1H-5
IIU
IIIL
M
R
1.14
1.71
0.53
0.81
0.33

-0.3
p 1.03
3.16
0.61
0.30
0.94
1.35
1.15
1.59
ST1H-6 IIIU M 1.25

r
0.8
e 0.36 1.96 0.37 1.07 1.30

P
ST1H-7 IV M 0.90 0.6 0.52 2.69 0.34 1.00 1.10

l
ST1H-8 VL M 1.20 0.53 0.45 1.35 0.56 1.04 1.12
ST1H-9
J94-1
VU
IL
M
R

n a
0.53
0.88
0.78
0.88
1.32
0.4
2.21
0.54
0.42
0.40
0.94
1.05
1.02
1.05
J94-2
J94-3
IU
IIL
u rR
M
1.29
0.88
0.53
0.58
0.61
0.65
0.83
0.98
0.96
0.68
0.83
0.98
0.83
0.98
J97-1
J97-2
J98-1
IL
IU
IIL Jo R
M
O
0.33
0.67
0.97
0.53
0.44
0.48
1.48
0.63
0.51
1.24
0.58
0.55
0.51
0.91
1.00
0.67
0.77
1.07
0.95
1.03
1.02
J116X-1 IL O 0.74 0.43 0.60 0.57 1.15 0.85 1.01
JG13-1 IU M 1.17 0.62 0.54 0.83 0.68 0.84 1.03

M- Marlstone, R- rudstone, O- Oil, Pr- Pristane, Ph- Phytane, TAR- Terrigenous/ aquatic ratio: (nC27 + nC29 + nC31 )/ (nC15 + nC17 + nC19 ) (Peters et al., 2005), CPI-
Carbon preference index: [(nC25 + nC27 + nC29 + nC31 + nC33 ) / (nC24 + nC26 + nC28+ nC30 + nC32 ) + (nC25 + nC27 + nC29+ nC31 + nC33 ) / (nC26 + nC28 + nC30 + nC32 +
i+1
nC34 )]/2 (Bray and Evans, 1961), OEP- Odd-to-even predominance: [(Ci + 6×Ci+2 +Ci+4 )/ (4×Ci+1 + 4×Ci+3 )] ×(-1) i=maxima carbon number (Marzi et al., 1993).

26
Journal Pre-proof

Table 2. Biomarker parameters of hopanes and steranes from the Shulu Sag.

Hopanes Steranes
Sample ID Unit Sample
Ts/(Tm+Ts) C30 M/C30 H Ga/C31 H C27 /C29 C28 /C29 C29 20S/(20S+20R) C29 ββ/(αα+ββ)

ST1H-1
ST1H-2
IL
IU
R
M
0.57
0.55
0.11
0.10
0.94
0.54
1.06
0.86
0.50
0.43
o f 0.45
0.46
0.55
0.55
ST1H-3 IIL M 0.51 0.09 0.29 0.99

r o
0.27 0.5 0.47
ST1H-4
ST1H-5
IIU
IIIL
M
R
0.38
0.54
0.11
0.13
0.08
0.08
0.67

-
0.47
p 0.23
0.25
0.45
0.42
0.42
0.4
ST1H-6 IIIU M 0.39 0.13

r
0.08
e0.50 0.20 0.41 0.41

P
ST1H-7 IV M 0.35 0.19 0.11 0.65 0.43 0.29 0.23

l
ST1H-8 VL M 0.33 0.20 0.11 1.01 0.28 0.32 0.18
ST1H-9
J94-1
VU
IL
M
R
0.39
0.71

n a
0.16
0.12
0.60
0.53
0.66
1.19
0.37
0.54
0.26
0.33
0.26
0.41
J94-2
J94-3
IU
IIL
R
M
0.34

u
0.38 r 0.16
0.17
0.60
0.21
0.92
0.79
0.42
0.32
0.27
0.30
0.27
0.23
J97-1
J97-2
J98-1
IL
IU
IIL
R
M
O J o
0.27
0.35
0.43
0.22
0.13
0.08
0.84
0.41
0.20
0.79
1.20
0.65
0.37
0.33
0.35
0.17
0.30
0.46
0.28
0.46
0.56
J116X-1 IL O 0.47 0.08 0.29 0.59 0.35 0.43 0.53
JG13-1 IU M 0.66 0.12 0.35 0.64 0.39 0.47 0.35

M- Marlstone, R- rudstone, O- Oil, C30 M- C30 Moretane, C30 H- C30 Hopane, Ga- Gammacerane, C31 H- C31 Hopane.

27
Journal Pre-proof

Fig. 1. Tectonic setting of the Shulu Sag, located in the southwestern corner of the Jizhong

Depression (I) of the Bohai Bay Bas in, which is composed of six depressions, denoted by letters

I-VI. (modified after Jiang et al., 2007).

Fig. 2. (A) Overview of the Shulu Sag. (B) Detailed structural map of the Shulu Sag. (C)

Stratigraphic system of the Shulu Sag.

L
Fig. 3. Correlation of well data and stratigraphic framework of the samples from the Es 3

submember. The well names with numbers (e.g., ST1H-1) indicate the location of the sample for

f
oo
biomarker analyses. Black arrows denote marlstone samples. Red arrows denote rudstone samples.

Green arrows denote oil samples from production.


pr
e-
Fig. 4. TOC contents of the nine different units in the Shulu Sag.
Pr

Fig. 5. Plot of average TOC contents versus average S 2 values of different units showing

hydrocarbon generation potential in the Shulu Sag (modified after Peters and Cassa, 1994).
al
rn

Fig. 6. (A) Plot of hydrogen index (HI) versus the pyrolysis temperature of maximum yield of

pyrolysate (Tma x) for different lithofacies within the Es 3 L submemeber in the Shulu Sag, displaying
u

the kerogen types and the thermal maturity stages (boundary lines modified from Mukhopadhyay
Jo

et al., 1995). (B) Plot of atomic H/C versus O/C.

Fig. 7. Gas chromatograms of saturated hydrocarbons (left column), mass chromatograms of

triterpanes (m/z=191; middle column), and steranes (m/z=217; right column) of selected samples
L
of Es 3 submemeber in the Shulu Sag. Locations of samples are shown in Fig. 3. C 30 H = C30

hopane; C30 M = C30 mortane; C27 RS = C27 regular steranes; C28 RS = C28 regular steranes; C29 RS

= C29 regular steranes. C30 Ster. = C30 regular steranes.

Fig. 8. (A) Plot of terrigenous/aquatic ratio (TAR) versus nC21 -/nC22 +. (B) Plot of

terrigenous/aquatic ratio (TAR) versus C27 /C29 sterane. R denotes hydrocarbons from rudsrone
28
Journal Pre-proof

samples and O indicates the oil samples obtained by production.

Fig. 9. (A) Plot of Pr/Ph versus Ga/C31 H. (B) Plot of Pr/nC17 versus Ph/nC18 (the boundary lines

modified from Shanmugam, 1985). Ga/C31 H = Gammacerane/C31 hopane. R denotes

hydrocarbons from rudsrone samples and O indicates the oil samples obtained by production.

Fig. 10. Ternary diagram of regular steranes (C27 , C28, and C29 ) from samples of Well ST1H

showing the relationship between sterane composition and organic matter sources (modified after

Huang and Meinschein, 1979).

f
oo
Fig. 11. Variations in vertical profile of Ro, Tma x, and maturity biomarkers of various units of Es 3L

in Well ST1H. C30 M/C30 H = C30 moretane/C30 hopane. pr


e-
Fig. 12. Variations in vertical profile of TOC, and source-related and environment-related
Pr

L
biomarkers of various units of Es 3 in Well ST1H. Ga/C31 H = Gammacerane/C31 hopane.
al

Fig. 13. (A) Plot of C30 M/C30 H versus Ts/(Ts+Tm). (B) Plot of C29 ββ/(αα+ββ) versus C29
rn

20S/(20S+20R). C30 M/C30 H = C30 moretane/C30 hopane.


u

Fig. 14. Hydrocarbon accumulation model of the Es 3 L submember in the Shulu Sag.
Jo

Fig. 15. Fracture characteristics in the Shulu Sag. (A) High-angle fractures with dissolution pores

(Unit I L from Well ST3: 4266.74 m). (B) Ruptured rudstones developed in a fault zone with

residual bitumen (Unit I L from Well ST3: 4273.6 m). (C) Horizontal fractures with residual

bitumen (Unit II L from Well J94: 3335.39 m). (D) Horizontal fractures with residual bitumen and

high-angle fractures (Unit II U from Well ST1H: 4074.53 m).

29
Journal Pre-proof

Declaration of interests

□√ The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

f
oo
pr
e-
Pr
al
u rn
Jo

30
Journal Pre-proof

Highlights

Aquatic organic matter has the best hydrocarbon generation potential

Three hydrocarbon accumulation pathways exist in the Shulu Sag

Biomarkers related to thermal maturity can indicate hydrocarbon migration

Source rock quality and migration ways influence oil reservoir heterogeneity

Mature, high TOC source rocks can form high-quality oil reservoirs

f
oo
pr
e-
Pr
al
u rn
Jo

31
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13
Figure 14
Figure 15

You might also like