You are on page 1of 11

Journal of Petroleum Science and Engineering 179 (2019) 696–706

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Organic matter origin and accumulation in tuffaceous shale of the lower T


Permian Lucaogou Formation, Jimsar Sag
Xiujian Dinga,b,∗, Jiangxiu Qua, Ablimit Iminc, Ming Zhaa, Yang Sua, Zhongfa Jianga, Hang Jiangd
a
School of Geosciences in China University of Petroleum, Qingdao, 266580, China
b
Shandong Provincial Key Laboratory of Depositional Mineralization & Sedimentary Minerals, Shandong University of Science and Technology, Qingdao, Shandong,
266590, China
c
Research Institute of Petroleum Exploration and Development, Xinjiang Petroleum Administration Bureau, Karamay, 841000, China
d
Strategic Research Center of Oil and Gas Resource, Ministry of Land and Resources, Beijing, 100034, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: The Jimsar Sag is an important shale oil exploration target area in the Junggar Basin, northwestern China. The
Tuffaceous shale lower Permian Lucaogou Formation, with a thickness of 200–300 m, is composed of thin tuffite, tuffaceous
Volcanic ash dolomite, mudstone and siltstone, and is the primary exploration target. A comprehensive analysis is conducted
Organic matter origin on the bulk and molecular geochemical characteristics of the tuffaceous shale, to define the organic matter
Organic matter accumulation
origin, depositional environment and organic matter accumulation. The key features of the tuffaceous shale
Lucaogou formation
extracts are short-to long-chain n-alkanes, a range of Pr/Ph ratio values (0.80–1.42), a relatively high content of
C29 steranes and a moderate content of gammacerane, showing that the redox state ranged from oxic to anoxic
and that the water salinity conditions varied from fresh to brackish. And the high content of cyanobacteria is the
primary cause of the high abundance of β-carotane and C29 sterane. It appears that volcanic ash did not change
the redox condition of depositional environment. Considering that volcanic ash-bearing sediments usually have
relatively high β-carotane and C29 sterane content, and previous studies have pointed out that volcanic ash can
stimulate phytoplankton and promote productivity in modern sediments, we could speculated volcanic ash may
release mineral ions such as iron, and then stimulated cyanobacteria, which was the main cause of organic
matter accumulation in the tuffaceous shale of the Jimsar Sag.

1. Introduction et al., 2011), such as the Green River model (Eugster and Surdam, 1973;
Burton et al., 2014), the anoxic lake model (Demaison and Moore,
Organic matter origin and accumulation in lacustrine source rocks 1980; Liang et al., 2018), the moderately deep lake model (Katz, 1990;
are essential to petroleum resource exploration and assessment (Kelts, Harris et al., 2004), and the small shallow lake model (Ding et al.,
1988; Katz, 1994; Hao et al., 2011), especially for shale oil and tight oil 2016). Tuffaceous shale components contain a certain amount of vol-
resources. Lacustrine source rock usually show strong vertical and canic ash, which has an obvious influence on water chemistry, organic
horizontal heterogeneity (Keym et al., 2006), and are still one of the matter productivity and degradation (Uematsu et al., 2004; Langmann
major uncertainties in petroleum exploration (Katz, 1990, 1995). et al., 2010; Turgeon and Creaser, 2008; Morel, 2003). However, we
Moreover, because of their sensitivity to climate changes, lacustrine found that research on organic matter origin and especially organic
source rocks have important scientific significance (Meyers, 2003; matter accumulation in tuffaceous shale was limited and unknown.
Adegoke et al., 2014). The Junggar Basin is a famous petroliferous basin in northwestern
Historically, productivity, organic matter preservation and dilution China, and tight oil and shale oil reserves are mainly distributed in the
have been the focus of studies on organic matter accumulation and Jimsar Sag (Qiu et al., 2016; Zou et al., 2013). Commercial oil flow has
petroleum source rock deposition (Arthur and Dean, 1998; Bohacs been obtained in the lower Permian Lucaogou Formation (LCG) from
et al., 2005). With different tectonic settings and depositional processes, nine exploration wells and the oil resource potential is mainly de-
there are many organic matter accumulation and source rock devel- termined by petroleum source rock (Kuang et al., 2013). Source rocks of
opment models of recent and ancient lake sediments (Katz, 2005; Hao Jimsar Sag are a set of thin lamellate tuffite, tuffaceous dolomite,


Corresponding author. No. 66, Changjiang West Road, Huangdao District, Qingdao, 266580 , China.
E-mail addresses: dingxj@upc.edu.cn, dingxj129@foxmail.com (X. Ding).

https://doi.org/10.1016/j.petrol.2019.05.004
Received 15 October 2018; Received in revised form 15 April 2019; Accepted 2 May 2019
Available online 06 May 2019
0920-4105/ © 2019 Elsevier B.V. All rights reserved.
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

mudstone and siltstone (Zhang et al., 2019). Although many studies


have analyzed tuffaceous sediments, the work has been focused on the
petrologic characteristics and reservoir properties, and not on the hy-
drocarbon source rock to study the organic matter origin and accu-
mulation (Grynberg et al., 1993; Ma et al., 2019). However, it has been
found that tuffaceous sediments have hydrocarbon-generating potential
in many areas of China. For example, studies on oil-source correlation
have demonstrated that crude oils distributed in the northwestern
Junggar Basin were mainly derived from organic-rich tuffaceous dolo-
mitic shales of the Fengcheng Formation (Zhou et al., 1989; Xiang et al.,
2015.). Tuffaceous shale of the Beisantai, Wucaiwan and Shanan areas
of the Eastern Junggar Basin, with average total organic carbon (TOC)
contents of 2.28%, 1.28% and 1.27%, respectively, were also proposed
as hydrocarbon source rocks (Wang et al., 2011). The present study
focuses on the organic matter origin and accumulation of tuffaceous
Fig. 1. Map of the Junggar Basin in northwestern China, showing the sample sediments.
well in the Jimsar Sag. Most source rock studies in the Jimsar Sag usually focused on pyr-
olysis methods and geochemical analysis (Kuang et al., 2012; Wei et al.,
2013; Peng et al., 2015; Cao et al., 2016), while organic matter origin
and accumulation of tuffaceous source rock were not a focus. A

Table 1
Total organic carbon and Rock-Eval pyrolysis data for tuffaceous sediments within the Lower Permian Lucaogou Formation in the Jimsar Sag, Junggar Basin.
Depth(m) Lithology TOC S1 S2 S1+S2 HI Tmax
(%) (mg/g) (mg/g) (mg/g) (S2/TOC*100) (°C)
3110.88 Tuffite 3.55 0.66 8.35 9.01 235.21 448
3112.09 Tuffaceous dolomite 0.85 0.42 1.92 2.34 225.88 443
3113.30 Tuffaceous dolomite 2.89 1.96 10.11 12.07 349.83 441
3113.34 Tuffaceous dolomite 0.72 0.02 0.71 0.73 98.61 443
3114.73 Tuffaceous dolomite 1.42 0.28 3.87 4.15 272.54 451
3118.78 Tuffite 9.77 0.44 78.96 79.40 808.19 453
3119.23 Tuffite 0.65 0.02 1.39 1.41 213.85 442
3134.05 Tuffaceous dolomite 6.77 0.74 32.75 33.49 483.75 444
3134.21 Tuffaceous dolomite 6.05 0.88 22.19 23.07 366.78 448
3144.66 Tuffaceous dolomite 0.55 0.06 0.99 1.05 180.00 442
3145.44 Tuffaceous dolomite 3.73 0.47 24.99 25.46 669.97 452
3146.16 Tuffaceous dolomite 2.72 0.79 6.98 7.77 256.62 441
3146.19 Tuffaceous dolomite 1.88 0.25 3.11 3.36 165.43 445
3152.82 Tuffite 3.17 0.02 20.20 20.22 637.22 451
3161.75 Tuffite 1.85 0.02 10.85 10.87 586.49 448
3162.02 Tuffite 3.22 0.17 10.76 10.93 334.16 453
3162.62 Tuffite 1.08 0.15 0.81 0.96 75.00 448
3174.75 Tuffite 0.43 0.03 0.79 0.82 183.72 447
3177.55 Tuffite 3.90 0.26 21.37 21.63 547.95 448
3177.57 Tuffite 3.39 0.21 7.80 8.01 230.09 455
3183.36 Tuffite 1.53 1.79 5.75 7.54 375.82 450
3183.80 Tuffite 3.85 0.22 7.85 8.07 203.90 452
3190.57 Tuffite 0.57 0.04 1.63 1.67 285.96 443
3200.73 Tuffaceous dolomite 0.53 0.04 2.03 2.07 383.02 449
3208.33 Tuffite 2.22 0.51 5.19 5.70 233.78 447
3212.16 Tuffite 1.13 0.21 1.62 1.83 143.36 449
3227.14 Tuffite 3.23 1.09 6.98 8.07 216.10 447
3263.19 Tuffaceous dolomite 1.21 0.04 1.79 1.83 147.93 444
3264.16 Tuffite 8.35 1.40 35.64 37.04 426.83 450
3274.00 Tuffaceous dolomite 0.45 0.01 0.69 0.70 153.33 451
3276.99 Tuffite 0.82 0.01 1.81 1.82 220.73 443
3277.65 Tuffite 6.83 2.42 28.82 31.24 421.96 445
3279.11 Tuffaceous dolomite 7.09 0.91 31.88 32.79 449.65 447
3283.74 Tuffaceous dolomite 0.61 0.02 1.02 1.04 167.21 442
3285.81 Tuffaceous dolomite 0.61 0.01 1.06 1.07 173.77 444
3286.70 Tuffaceous dolomite 8.16 1.16 39.94 41.10 489.46 446
3305.33 Tuffite 0.65 0.03 1.42 1.45 218.46 449
3306.47 Tuffite 0.98 0.36 2.06 2.42 210.20 441
3306.97 Tuffite 0.65 0.09 0.87 0.96 133.85 441
3308.42 Tuffaceous dolomite 0.56 0.03 0.74 0.77 132.14 439
3309.40 Tuffaceous dolomite 3.76 0.39 16.24 16.63 431.91 453
3311.18 Tuffite 0.72 0.26 1.68 1.94 233.33 438
3312.59 Tuffite 0.76 0.50 0.36 0.86 47.37 441
3313.65 Tuffite 2.20 0.27 3.06 3.33 139.09 450
3319.96 Tuffaceous dolomite 1.62 0.05 4.55 4.60 280.86 448
3323.38 Tuffaceous dolomite 4.72 1.46 18.91 20.37 400.64 449

Note: TOC = Total organic carbon (wt. %); S1 (mg/g), the amount of hydrocarbon of free; S2 (mg/g), the amount of hydrocarbon produced by the cracking of organic
matter; Tmax (°C), the temperature which corresponds to the maximum generation rate from kerogen cracking.

697
Table 2
Biomarker parameters for tuffaceous sediments within the Lower Permian Lucaogou Formation in the Jimsar Sag, Junggar Basin.
X. Ding, et al.

Depth(m) Lithology Isoprenoid ratios Alkanes Steranes Terpanes

Pr/Ph Pr/nC17 Ph/nC18 CPI C21-/C22+ 27% 28% 29% C27/(C27+C29) C29/C27 20S(%) ββ(%) G/H C22T/C21T C24T/C23T C26T/C25T C24Te/C26T

3110.88 Tuffite 1.08 2.67 2.43 1.17 0.53 15 33 52 0.22 3.46 0.32 0.23 0.22 0.13 0.42 2.70 1.12
3112.09 Tuffaceous dolomite 1.17 1.17 1.11 1.21 0.85 21 31 48 0.30 2.34 0.39 0.26 0.22 0.13 0.44 2.46 1.07
3113.30 Tuffaceous dolomite 1.22 1.09 1.10 0.83 1.11 19 33 48 0.28 2.53 0.35 0.22 0.36 0.15 0.34 2.11 1.46
3113.34 Tuffaceous dolomite 1.12 0.94 0.85 1.19 0.69 16 35 49 0.24 3.16 0.46 0.36 0.16 0.16 0.45 2.21 1.14
3114.73 Tuffaceous dolomite 1.21 1.21 1.18 1.32 0.57 32 29 39 0.45 1.24 0.36 0.21 0.17 0.11 0.40 2.40 1.37
3118.78 Tuffite 1.51 0.63 0.58 0.96 2.02 18 34 48 0.27 2.73 0.41 0.26 0.22 0.13 0.38 2.26 2.15
3119.23 Tuffite 1.18 0.90 1.80 1.18 0.87 15 34 51 0.23 3.40 0.45 0.35 0.16 0.15 0.41 2.22 1.18
3134.05 Tuffaceous dolomite 1.42 1.89 1.73 0.66 0.59 23 32 45 0.34 1.98 0.32 0.19 0.18 0.11 0.39 2.67 1.38
3134.21 Tuffaceous dolomite 1.24 2.02 1.62 1.65 0.63 23 31 47 0.33 2.05 0.31 0.18 0.26 0.12 0.41 2.96 1.35
3144.66 Tuffaceous dolomite 1.15 1.31 1.34 1.20 0.87 13 35 52 0.21 3.85 0.43 0.29 0.18 0.14 0.42 2.43 1.11
3145.44 Tuffaceous dolomite 1.16 0.87 0.86 1.11 1.03 16 37 47 0.26 2.86 0.40 0.26 0.18 0.13 0.36 2.08 2.22
3146.16 Tuffaceous dolomite 1.2 0.54 0.38 1.12 1.70 19 41 40 0.32 2.16 0.30 0.18 0.07 0.15 0.33 2.17 1.65
3146.19 Tuffaceous dolomite 1.03 0.5 0.45 1.25 0.93 19 45 36 0.35 1.88 0.32 0.17 0.09 0.12 0.44 2.44 1.11
3152.82 Tuffite 1.26 0.88 0.71 1.27 0.58 15 54 31 0.32 2.10 0.41 0.26 0.13 0.14 0.42 2.16 2.00
3161.75 Tuffite 1.49 0.17 0.08 1.34 0.91 20 42 38 0.35 1.84 0.32 0.20 0.14 0.13 0.44 2.41 4.29
3162.02 Tuffite 1.46 0.15 0.08 1.25 0.99 19 42 40 0.32 2.12 0.32 0.18 0.15 0.28 0.52 2.06 7.27
3162.62 Tuffite 1.50 0.15 0.09 1.21 0.64 22 27 51 0.30 2.29 0.35 0.19 0.07 0.23 0.53 2.04 12.07
3174.75 Tuffite 1.26 0.36 0.23 1.36 0.68 14 45 41 0.26 2.87 0.36 0.19 0.16 0.13 0.40 2.31 1.92
3177.55 Tuffite 1.07 0.25 0.22 1.11 0.49 22 38 40 0.36 1.81 0.37 0.21 0.09 0.17 0.43 2.18 3.62
3177.57 Tuffite 1.05 0.26 0.22 1.24 0.51 22 37 41 0.35 1.90 0.37 0.21 0.09 0.17 0.43 2.06 3.67
3183.36 Tuffite 1.4 0.3 0.17 1.27 0.65 19 37 44 0.30 2.29 0.31 0.18 0.15 0.14 0.39 2.22 1.74
3183.80 Tuffite 1.35 0.35 0.22 1.15 0.79 20 39 41 0.33 2.01 0.30 0.16 0.18 0.13 0.36 2.10 2.59
3190.57 Tuffite 1.13 1.34 1.42 1.19 0.88 13 35 53 0.19 4.22 0.45 0.32 0.18 0.14 0.42 2.24 1.14

698
3200.73 Tuffaceous dolomite 1.42 1.58 1.26 1.25 0.86 18 34 48 0.27 2.67 0.40 0.24 0.19 0.13 0.39 2.68 1.05
3208.33 Tuffite 1.42 1.63 1.27 0.98 0.66 18 36 46 0.29 2.47 0.35 0.19 0.25 0.16 0.35 1.96 2.17
3212.16 Tuffite 0.94 0.92 1.18 1.47 0.39 17 37 46 0.27 2.72 0.42 0.26 0.13 0.20 0.75 3.62 10.22
3227.14 Tuffite 1.13 1.55 1.86 1.21 0.83 15 37 48 0.24 3.25 0.38 0.21 0.19 0.11 0.39 2.72 1.55
3263.19 Tuffaceous dolomite 1.06 1.75 2.40 1.15 1.08 12 35 53 0.18 4.44 0.43 0.27 0.19 0.14 0.43 2.50 1.13
3264.16 Tuffite 1.19 1.27 1.68 1.03 2.01 14 33 53 0.21 3.68 0.44 0.25 0.21 0.12 0.38 2.11 1.91
3274.00 Tuffaceous dolomite 1.05 1.77 2.35 1.16 0.95 12 33 55 0.18 4.62 0.44 0.28 0.19 0.13 0.42 2.50 1.11
3276.99 Tuffite 1.10 1.75 2.26 1.14 1.07 11 34 55 0.17 4.75 0.44 0.29 0.20 0.14 0.40 2.54 1.11
3277.65 Tuffite 1.11 1.53 2.04 1.40 1.33 12 34 54 0.19 4.30 0.42 0.25 0.20 0.13 0.41 2.40 1.35
3279.11 Tuffaceous dolomite 0.99 1.78 2.55 1.41 0.93 11 34 55 0.17 4.85 0.41 0.24 0.19 0.12 0.42 2.41 1.03
3283.74 Tuffaceous dolomite 1.03 1.82 2.59 1.13 1.10 11 33 56 0.17 4.97 0.43 0.25 0.19 0.13 0.44 2.54 0.99
3285.81 Tuffaceous dolomite 1.03 1.81 2.48 1.10 1.23 11 32 57 0.16 5.10 0.42 0.25 0.18 0.13 0.42 2.66 0.97
3286.70 Tuffaceous dolomite 1.01 1.49 2.17 1.21 1.44 14 32 54 0.21 3.71 0.43 0.25 0.20 0.10 0.41 2.18 1.52
3305.33 Tuffite 0.80 1.65 2.73 1.24 0.61 15 33 52 0.22 3.47 0.42 0.24 0.15 0.12 0.45 2.85 0.57
3306.47 Tuffite 0.85 1.51 2.62 1.22 1.14 15 34 51 0.22 3.48 0.40 0.23 0.15 0.13 0.44 2.97 0.56
3306.97 Tuffite 0.87 1.53 2.58 1.23 1.13 15 33 52 0.22 3.53 0.41 0.24 0.15 0.13 0.42 2.70 0.57
3308.42 Tuffaceous dolomite 0.88 1.48 2.22 1.31 0.56 13 41 47 0.21 3.68 0.41 0.24 0.14 0.14 0.42 2.78 0.66
3309.40 Tuffaceous dolomite 0.84 1.52 2.66 1.50 0.85 12 35 53 0.19 4.39 0.41 0.22 0.13 0.14 0.41 2.70 0.72
3311.18 Tuffite 1.03 1.69 2.36 1.42 1.06 17 37 46 0.28 2.62 0.43 0.25 0.16 0.13 0.45 3.09 0.72
3312.59 Tuffite 0.96 1.63 2.30 1.18 0.85 17 35 48 0.26 2.80 0.42 0.25 0.14 0.11 0.45 3.17 0.76
3313.65 Tuffite 0.99 1.88 2.50 1.36 0.74 9 30 61 0.13 7.00 0.39 0.34 0.21 0.19 0.41 2.73 1.62
3319.96 Tuffaceous dolomite 1.01 1.24 1.74 1.46 1.80 18 33 49 0.27 2.64 0.42 0.26 0.19 0.14 0.41 2.69 0.64
3323.38 Tuffaceous dolomite 0.95 1.38 1.73 1.34 0.88 18 34 48 0.27 2.65 0.42 0.29 0.20 0.15 0.42 2.70 0.65

Note: Pr/Ph = Pristane/Phytane; Pr/C17 = Pristane/C17 n-alkane; Ph/C18 = Phytane/C18 n-alkane; CPI = 2(C23+C25+C27+C29)/[C22+2(C24+C26+C28)+C30]; %27 = C27 steranes/(C27 steranes + C28 steranes + C29
steranes); %28 = C28 steranes/(C27 steranes + C28 steranes + C29 steranes); %29 = C29 steranes/(C27 steranes + C28 steranes + C29 steranes); C27/(C27+C29) = C27 steranes/(C27 steranes + C29 steranes); C29/
C27 = C29 steranes/C27 steranes; 20S(%) = C29 sterane ααα20S/(20S + 20R); ββ(%) = C29 sterane αββ/(αββ+ααα); G/H = Gammacerane/C30 hopane; C22T/C21T = C22 tricyclic terpane/C21 tricyclic terpane; C24T/
C23T = C24 tricyclic terpane/C23 tricyclic terpane; C26T/C25T = C26 tricyclic terpane/C25 tricyclic terpane; C24Te/C26T = C24 tetracyclic terpane/C25 tricyclic terpane.
Journal of Petroleum Science and Engineering 179 (2019) 696–706
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

Fig. 2. Vertical variations in TOC, S1, HI and Tmax data for tuffaceous sediments within the Lower Permian Lucaogou Formation in the Jimsar Sag of the Junggar
Basin. TOC = Total organic carbon (wt. %); S1 (mg/g), the amount of hydrocarbon of free; HI = S2/TOC*100, S2 (mg/g), the amount of generated hydrocarbons;
Tmax (°C), the temperature of the maximum generation rate of kerogen cracking.

comprehensive study of geochemical characteristics has been con- boundary faulting in the west, and has been a main exploration target
ducted on the tuffaceous shale of the LCG in the Jimsar Sag, to (1) of unconventional oil exploration in China (Wang et al., 2015; Jiang
reveal geochemical characteristics and hydrocarbon generation poten- et al., 2015; Cao et al., 2017).
tial; (2) investigate depositional environment and organic matter With a Carboniferous basement, the sediments were deposited from
origin; and (3) define controlling factors on organic matter accumula- the Permian to Quaternary. Tight oil is mainly contained in the LCG
tion and present a tuffaceous source rock development model. (Cao et al., 2017). During deposition of the LCG, volcanic material,
terrigenous detrital materials and carbonate were mixed and caused
typical mixed sediments (Shao et al., 2015). Comprehensive application
2. Geological setting of core and thin section observations show that the LCG, with a
thickness of 200–300 m, is composed of tuffite, tuffaceous dolomite,
The Junggar Basin is a famous petroliferous basin in northwestern mudstone and siltstone and could be further subdivided into the lower
China, and the Jimsar Sag is an important sag with proven tight oil and (P2l1) and upper Lucaogou Member (P2l2) (Jiang et al., 2015). The
shale oil resources, located in the eastern basin (Fig. 1) (Ding et al., single tuffite and tuffaceous dolomite layers are generally 2.0–3.0 m
2017; Li et al., 2019). The Jimsar Sag is a small half-graben sag with

699
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

be affected. And the Tmax data may be uncertain ( ± 10 °C).


Gas chromatography-mass spectrometry (GC-MS) analysis was per-
formed by an HP680GC/5973MSD instrument equipped with an HP-
5MS fused silica column (30 m × 0.25 mm i.d., film thickness 0.25 μm).
First, the organic matter was conducted by chloroform/methanol
(87:13, v/v) and separated into several parts: saturated hydrocarbon,
aromatic hydrocarbon, and asphaltine. Second, the saturated hydro-
carbon was dissolved in hexane and analyzed by an HP-5MS column.
The analysis results are listed in Table 2.
Polished whole-rock blocks were prepared for organic matter clas-
sification via a Zeiss Axio Imager fluorescence microscope.
Characteristics of the macerals were analyzed under reflected white
light and incident blue light. According to the well-established ICCP
liptinite classification (Pickel et al., 2017a,b), the visual analysis reveals
mainly liptinite, but areas of vitrinite and inertinite were also present.

4. Results
Fig. 3. Total organic carbon (TOC) contents versus Rock-Eval S2 peaks (mg
hydrocarbons [HC]/g rock) for tuffaceous sediment samples from the Lucaogou 4.1. Source rock quality
Formation (P2l) of the Jimsar Sag.
TOC, S2, and HI (hydrogen indices, S2/TOC) are usually used to
evaluate organic matter abundance, and kerogen types (Tissot and
Welte, 1984). The TOC of the LCG tuffaceous source rocks ranges from
0.43% to 9.77%, with a mean value of 2.66% (Table 1, Fig. 2). The
hydrocarbon generation index (S1+S2) ranges between 0.70 and
41.10 mg HC/g·TOC, with mean value of 11.17 mg HC/g·TOC (Table 1,
Fig. 2). Generally, HI values higher than 6.0 mg HC/g represent higher
hydrocarbon potential (Tissot and Welte, 1984). Cross plots of S2 and
TOC for the LCG tuffaceous samples are shown in Fig. 3. Most of the
LCG samples had TOC higher than 1.0% and S2 higher than 2 mg HC/g
rock, which suggested fair, good and excellent source rocks were de-
veloped in the sag, based on the guidelines from Peters and Cassa
(1994).
Rock-Eval pyrolysis parameters are also usually used to classify
organic matter type (Peters, 1986). LCG source rock samples show wide
variations in kerogen types, according to the wide range of HI
(47.37–808.19 mg/g TOC) (Fig. 4). According to the organic matter
types standard (Peters, 1986; Huang et al., 1984), the kerogen com-
position of the LCG tuffaceous samples is mainly distributed from type I
to type II, with a small portion of type III (Fig. 4). And type II may be
just a mixing of type I and type III.
The Tmax values of LCG tuffaceous samples range from 438 to
455 °C, implying that the thermal evolution is between the immature
and mature stages (Table 1, Fig. 2), which is consistent with the Ro
values (0.7–1.0) in Pang (2019).

4.2. Organic matter visual analysis


Fig. 4. Plot of hydrogen index versus Tmax outlining the kerogen type of
Lucaogou Formation in the Jimsar Sag of the Junggar Basin.
The Liptinite group includes bituminite, alginate, cutinite, sporinite,
suberinite, resinite, exsudatinite and liptodetrinite (Pickel et al.,
thick and strongly heterogeneous. The LCG consist of mainly shallow 2017a,b). In tuffaceous shale of the Lucaogou Formation studied here,
and semi deep lake facies, with minor deep lake facies (Cao et al., alginite was observed in relatively high abundances. Alginite was ob-
2017). served as locally agglomerated (Fig. 5a), stripe-like and parallel to
bedding (Fig. 5b and c). Cutinite usually showed yellow fluorescence
3. Samples and methods (Fig. 5d) and vitrinite did not have any fluorescence in the samples of
tuffaceous shale (Fig. 5e). Inertinite showed a higher reflection than
A total of 46 core samples were collected from the LCG tuffaceous vitrinite under reflected white light, and no fluorescence under incident
shale cores in the Jimsar Sag. The sampling well is located in the central blue light illumination (Fig. 5f).
part of the sag and is shown in Fig. 1B. The 46 core samples were used
for TOC content measurements, pyrolysis analysis and gas chromato- 4.3. Molecular analysis
graphy-mass spectrometry (GC-MS) analysis.
To obtain the TOC and pyrolysis parameters (Tmax, S1 and S2), a Biomarkers are often used to assess organic matter origin, deposi-
Rock-Eval VI instrument was used and the procedure was described by tional environment (e.g. redox condition and salinity) and thermal
Behar et al. (2001). The data are shown in Table 1. It should be men- maturity stage, and have drawn increasing attention in recent years
tioned that occurrence of hydrocarbons in samples will always affect (Peters et al., 2005; Arfaoui et al., 2007). The biomarker distributions of
the shape of the S2 peak in the FID and reading of the Tmax value may selected LCG tuffaceous shale samples are shown in Fig. 6.

700
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

Fig. 5. Photomicrographs of organic matter in the polished blocks: a-c, alginite under incident blue light; d, cutinite under incident blue light; e, vitrinite under
incident blue light; f, inertinite under reflected white light. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web
version of this article.)

4.3.1. N-alkanes and isoprenoids of the C27/(C27 + C29) sterane ratio are high in the upper LCG
The tuffaceous shale extracts contain a full range of n-alkanes and (0.21–0.47), whereas lower values are observed in the lower LCG
isoprenoids. The patterns of n-alkane are dominated by short-to long- (0.12–0.28) (Table 2).
chain n-alkanes (Fig. 6). The values of carbon preference index (CPI)
range from 0.65 to 1.65, and the C21-/C22+ values range from 0.39 to 5. Discussion
2.02, as shown in Table 2. The Pr/Ph values of the upper LCG are higher
than those of the lower LCG, ranging from 1.03 to 1.51 and 0.80 to 5.1. Origin of organic matter
1.42, respectively (Table 2). The extracts also display a high content of
β-carotane. The height of β-carotane is as high as the major n-alkane N-alkane distribution patterns are also usually used to define or-
peak, or even higher (Fig. 6). ganic matter origin, e.g. Peters et al. (2005) reported that source rocks
with CPI∼1 may show that the organic matter origin was phyto-
4.3.2. Terpanes plankton, zooplankton and benthic bacteria. The mean CPI value of all
The tricyclic and pentacyclic terpane distribution pattern obtained the samples is approximately 1.22, reflecting a major contribution of
from m/z 191 ion chromatograms is shown in Fig. 6. Gam/C30H (G/H) phytoplankton and aquatic algal-bacterial organisms. Long-chain n-al-
values range from 0.07 to 0.36 (Table 2). Both Tm and Ts are present in kanes (C22–C35) chiefly come from terrestrial higher plants and short-
all samples, and Tm is significantly higher than Ts, as shown in Fig. 2 chain n-alkanes (C12–C21) are typical of algae and plankton (Tissot and
and Table 2. Tricyclic terpane abundances are relatively lower (Fig. 6), Welte, 1984; Eglinton and Hamilton, 1967; Cranwell et al., 1987;
and C21T is generally greater than C19T and C23T (Table 2). Meyers, 1997). The values of NC21-/NC22+ mainly range from 0.5 to 1.5
(Fig. 7), indicating that the organic matter of LCG was probably mainly
4.3.3. Steranes derived from an algal or planktonic source with terrestrial plant input
The sterane distribution patterns are characterized by a dominant (see Fig. 8).
regular steranes, and relatively low abundances of diacholestanes Proportional sterane abundances are usually used as useful in-
(Fig. 6). Stigmastane (C29) is generally greater than ergostane (C28) and dicators for defining organic matter biological sources (Seifert and
cholestane (C27) among the regular steranes in the LCG (Fig. 6). Values Moldowan, 1978; Huang and Meinschein, 1979). The ternary diagram

701
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

Fig. 6. Biomarker distribution of selected tuffaceous samples of Lucaogou Formation in the Jimsar Sag of the Junggar Basin. Show in the sequence of TIC, sterane (m/
z 217) and terpane (m/z 191) distribution. Pr = pristane; Ph = phytane; C19T = C19 tricyclic terpane; C21T = C21 tricyclic terpane; C23T = C23 tricyclic terpane;
C24Te = C24 tetracyclic terpane; Ts = 18α (H)-22, 29, 30-trisnorneohopane; Tm = 17α (H)-22, 29, 30-trisnorhopane; C30H = C30 hopane; C27 = C27 sterane 20R;
C28 = C28 sterane 20R; C29 = C29 sterane 20R.

of C27–C28–C29 steranes of the LCG tuffaceous sediment samples is because of water column stratification caused by volcanic ash (Meyer
shown in Fig. 7. The abundant steranes in the LCG tuffaceous sediment and Kump, 2008; Sinton and Duncan, 1997).
samples are C29 (27%–60%, average 47%), followed by C28 (31%–59%, The Pr/Ph (pristane/phytane) ratio is an indicator redox condition
average 37%) with minimal C27 steranes (8%–33%, average 17%), in a depositional environment. It is usually considered that a ratio less
which is shown in Table 2 and Fig. 7. than 1 may imply a suboxic or anoxic environment, while a ratio
It is usually considered that C27 steranes are derived mainly from greater than 1 may indicate oxidizing depositional conditions (Didyk
phytoplankton (e.g., Volkman, 1983) and that C28 steranes are derived et al., 1978). It should be mentioned that some researchers believe the
from specific phytoplankton types, such as diatoms containing chlor- ratio may be influenced by thermal maturity, and cannot be used as an
ophyll-c (Knoll et al., 2007), whereas C29 steranes mainly originate indicator of depositional redox conditions (Connan, 1973; Albrecht
from terrigenous higher plants (e.g., Hao et al., 2011; Volkman, 2003). et al., 1976). In our study area, the ratios of tuffaceous shale decrease
In general, the steranes of the tuffaceous shale of P2l may reflect a low with increasing burial depth (Fig. 6). Therefore, it could be concluded
contribution of phytoplankton and aquatic algal-bacterial organisms that the ratio is not controlled by thermal maturity. The Pr/Ph ratios
and a major contribution of terrestrial higher plants. However, it should range from to 0.8 to 1.5, with a mean value of 1.1, implying a relatively
be noted that bituminitel and alginite were observed with relatively mildly oxic to mildly anoxic depositional environment.
high contents in the samples, the δ13C of the kerogen ranges from Plots of the Pr/NC17 ratio versus Ph/NC18 ratio shown in Fig. 9
−28.33‰ to −24.28‰ (average −26.69‰) (Wu et al., 2017), and could also be used to indicate redox conditions of the organic matter
cyanobacteria are also an important source of C29 steranes (Volkman, depositional environment (Peters et al., 1999), indicating a mildly
1983; Volkman et al., 1998). Thus, it appears that a high portion of the oxidizing to mildly reducing depositional environment. The tuffaceous
C29 steranes are associated with cyanobacteria, as the main source of sediment depositional environment of the lower LCG is more reducing
organic matter. Considering that volcanic ash can fertilize surface wa- than that of the upper LCG (Fig. 9). It appears that volcanic ash did not
ters and stimulate phytoplankton (Duggen et al., 2007; Kurenkov, change the redox conditions in the P2l tuffaceous sediment in the
1966). Cyanobacteria is rich in carotenoids, and is also the expected Jimsar Sag.
precursor of β-carotane (Philp et al., 1992). Therefore, we concluded Gammacerane is derived from the reduction of tetrahymanol
that the high content of cyanobacteria is also the primary cause of the (Venkatesan, 1989), and the principal source of tetrahymanol appears
high abundance of β-carotane in the LCG tuffaceous sediment samples. to be bacterivorous ciliates, which usually stay at the interface between
oxic and anoxic zones of stratified water columns (Sinninghe et al.,
1995). Therefore, gammacerane could be used to imply a stratified
5.2. Depositional environment water column. Although a stratified water column can also be attrib-
uted to either hypersalinity or temperature gradient (e.g., Bohacs et al.,
Redox conditions in the depositional environment are the essential 2000), a high content of gammacerane is mostly found in evaporite or
controlling factors for organic matter degradation and preservation high-salinity depositional environments (Holba et al., 2003; Summons
(Burdige, 2007). Previous studies have indicated that volcanic ash has et al., 2008). The gammacerane index (gammacerane/C30hopane)
contributed to forming an anoxic environment in modern and ancient ranges from 0.07 to 0.36 (average 0.17) and indicates a fresh to
sediments (Jones and Gislason, 2008; Turgeon and Creaser, 2008). A brackish water depositional environment (Hao et al., 2011), which is
very strong line of scientific evidence is that the temporal and spatial shown in Table 2 and Fig. 7. It should be mentioned that the brackish
distribution of the Cretaceous oceanic anoxic event is consistent with a water depositional environment may also be caused by volcanic ash
volcano eruption (Turgeon and Creaser, 2008), and mainly formed

702
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

Fig. 7. Vertical variations in TOC, Pr/Ph, C21-/C22+, G/H, 27%, 28% and 29% data for tuffaceous sediments within the Lower Permian Lucaogou Formation in the
Jimsar Sag of the Junggar Basin. Abbreviations for biomarker parameters are defined in Table 2.

(Censi et al., 2010), as the gammacerane index is relatively higher in


the condensed tuffaceous shale.

5.3. Organic matter accumulation

Organic matter accumulation is a very complex process that is


controlled by many factors, such as climate, water chemistry, weath-
ering, and tectonic setting. Historically, organic matter productivity,
preservation, and sedimentation rate have been considered as the three
main controlling factors of organic matter accumulation (Bohacs et al.,
2005; Calvert, 1990; Arthur and Dean, 1998; Katz, 2005). In the Jimsar
Sag, tectonic movement was very weak during deposition of the Per-
mian LCG, and the sedimentation rate of the LCG was relatively steady,
as proven by the less variable strata thickness (Kuang et al., 2013). The
volcanic ash was transported to the lake mainly by wind and had the
little effect on sedimentation rate (Yang et al., 2018). Therefore, we
conclude that the sedimentation rate is not the key controlling factor of
organic matter accumulation.
Volcanic ash can fertilize surface water and enhance primary pro-
ductivity by releasing mineral ions such as iron in modern lakes and
oceans (Kurenkov, 1966; Uematsu et al., 2004; Langmann et al., 2010).
Fig. 8. Ternary diagram of C27–C28–C29 steranes for tuffaceous sediments
Duggen et al. (2007) recognized that volcanic ash could stimulate
within the Lower Permian Lucaogou Formation in the Jimsar Sag of the Junggar
phytoplankton and promote productivity-based biogeochemical ex-
Basin (Modified after Adegoke et al., 2014). Abbreviations for biomarker
periments. Biomarkers, especially sterane parameters have the poten-
parameters are defined in Table 2.
tial to indicate primary producers (e.g., Volkman et al., 1998; Hao et al.,

703
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

drawn:

(1) The tuffaceous shale is a good source rocks with high hydrocarbon
generation potential, as indicated by high TOC (total organic
carbon, 0.43%–9.77%) and S1+S2 (hydrocarbon potential index
0.70–41.10 mg HC/g).
(2) The molecular composition of the Lucaogou tuffaceous sample ex-
tracts reflects fresh to brackish water conditions and a relatively
mildly oxidizing to mildly reducing depositional environment. It
appears that volcanic ash did not change the redox condition but
caused the brackish water depositional environment.
(3) Volcanic ash may have released mineral ions and then stimulated
cyanobacteria blooming, which was associated with the high por-
tion of C29 steranes and the main reason for organic matter accu-
mulation in the tuffaceous sediment of the Jimsar Sag.

Acknowledgments

Fig. 9. Relationship between Ph/nC18 and Pr/nC17 ratio values for tuffaceous
This work was supported by a grant from the National Science
sediments within the Lower Permian Lucaogou Formation in the Jimsar Sag of
the Junggar Basin (Modified after Peter et al., 1999). Abbreviations for bio-
Foundation for Young Scientists of China (Grant No. 41702143), the
marker parameters are defined in Table 2. Fundamental Research Funds for the Central Universities
(17CX02006A), and the Foundation of Shandong Provincial Key
Laboratory of Depositional Mineralization & Sedimentary Mineral
2011).
(DMSM2017063). We would like to thank the Xinjiang Petroleum
To define whether productivity had a controlling effect on organic
Administration Bureau of CNPC for providing the core samples. Special
matter accumulation, we comprehensively analyzed the relationship
thanks to Executive Editor Tahar Aifa, reviewer Vincent Crombez and
between TOC and sterane content in the P2l tuffaceous sediment sam-
another anonymous reviewer for their constructive comments and
ples. The TOC of the LCG tuffaceous shale is characterized by high
corrections, which have substantially improved the quality of this
values and strong lateral heterogeneity, with a range of 0.43%–9.77%
manuscript.
(Fig. 7). The C27 and C28 sterane content tendency are obviously dif-
ferent from the TOC content tendency, while the C29 sterane content
Appendix A. Supplementary data
displays the same trend as the TOC content, which increasing and de-
creasing in the lower and upper LCG, separately. Therefore, we can
Supplementary data to this article can be found online at https://
conclude that organic matter accumulation is controlled by the abun-
doi.org/10.1016/j.petrol.2019.05.004.
dance of cyanobacteria (the major source of C29 steranes) during LCG
deposition. Moreover, it should also be mentioned that the gamma-
References
cerane index (gammacerane/αβ C30 hopane) has the same change trend
as TOC and C29 steranes, indicating that water salinity had an apparent
Adegoke, A.K., Abdullah, W.H., Hakimi, M.H., et al., 2014. Geochemical characterisation
influence on organic matter accumulation. of Fika Formation in the Chad (Bornu) Basin, northeastern Nigeria: implications for
In addition, redox conditions are commonly considered to be the depositional environment and tectonic setting. Appl. Geochem. 43, 1–12.
Albrecht, P., Vandenbroucke, M., Mandengue, M., 1976. Geochemical studies on the
key factor for organic matter degradation and preservation (Hedges and
organic matter from the Douala Basin (Cameroon)-I. Evolution of the extractable
Keil, 1995; Demaison and Moore, 1980). The Pr/Ph displays no re- organic matter and the formation of petroleum. Geochem. Cosmochim. Acta 40,
lationship to the TOC. Although the redox conditions of the lower and 791–799.
upper LCG are obviously different, the TOC content values are similar. Arfaoui, A., Montacer, M., Kamoun, F., et al., 2007. Comparative study between Rock-
Eval pyrolysis and biomarker parameters: a case study of Ypresian source rocks in
Thus, the implied redox conditions did not have evident controlling on central-northern Tunisia. Mar. Petrol. Geol. 24, 566–578.
organic matter accumulation. Arthur, M.A., Dean, W.E., 1998. Organic matter production and preservation and evo-
Based on a comprehensive analysis of organic productivity, organic lution of anoxia in the Holocene Black Sea. Paleoceanography 13, 395–411.
Behar, F., Beaumont, V., Penteado, H.D.B., 2001. Rock-Eval 6 technology: performances
matter preservation and accumulation, we concluded productivity is and developments. Oil Gas Sci. Technol. 56 (2), 111–134.
the main controlling factor of organic matter accumulation. Previous Bohacs, K.M., Carroll, A.R., Neal, J.E., et al., 2000. Lake-basin type, source potential, and
studies had proved that volcanic ash can stimulate phytoplankton and hydrocarbon character. An integrated sequence-stratigraphic geochemical frame-
work. In: Gierlowski-Kordesch, E.H., Kelts, K.R. (Eds.), Lake Basins through Space
promote productivity in modern sediments (Duggen et al., 2007; and Time. American Association of Petroleum Geologists Studies in Geology. vol. 46.
Uematsu et al., 2004; Langmann et al., 2010). Therefore, we could pp. 3–34.
conclude volcanic ash was the crucial ingredient on causing high pro- Bohacs, K.M., Grabowski Jr., G.J., Carroll, A.R., et al., 2005. Productivity, destruction,
and dilution-the many paths to source-rock development. In: Harris, N.B. (Ed.), The
ductivity in LCG tuffaceous sediments, which is supported by that ash-
Deposition of Organic-Carbon-Rich Sediments: Models, Mechanisms, and
bearing sediments usually have relatively high β-carotane and C29 Consequences. vol. 82. Society for Sedimentary Geology Special Publication, pp.
sterane content. We speculated volcanic ash may release mineral ions 61–101.
Burdige, D.J., 2007. Preservation of organic matter in marine sediments: controls, me-
such as iron, and then stimulated cyanobacteria, which was the main
chanisms, and an imbalance in sediment organic carbon budgets? Chem. Rev. 107
cause of organic matter accumulation in the tuffaceous shale of the (2), 467–485.
Jimsar Sag. Burton, D., Woolf, K., Sullivan, B., 2014. Lacustrine depositional environments in the
Green River formation, uinta basin: expression in outcrop and wireline logs. AAPG
(Am. Assoc. Pet. Geol.) Bull. 98, 1699–1715.
6. Conclusions Calvert, T.F., 1990. Anoxia vs. productivity: what controls the formation of organic
carbon rich sediments and sedimentary rocks? AAPG (Am. Assoc. Pet. Geol.) Bull. 4,
454–466.
Based on our bulk geochemical study and molecular geochemical Cao, Z., liu, G., Kong, Y., et al., 2016. Lacustrine tight oil accumulation characteristics:
observations of 47 tuffaceous shale samples, we inferred the organic permian Lucaogou Formation in jimusaer sag, Junggar Basin. Int. J. Coal Geol. 153,
37–51.
matter origin, depositional environment and organic matter accumu-
Cao, Z., Liu, G., Xiang, B., et al., 2017. Geochemical characteristics of crude oil from a
lation of the tuffaceous shale, and the following conclusions could be

704
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

tight oil reservoir in the Lucaogou Formation, Jimusar sag, Junggar Basin. AAPG 154.
(Am. Assoc. Pet. Geol.) Bull. 101, 39–72. Ma, J., Huang, Z., Li, T., 2019. Mechanism of hydrocarbon accumulation and enrichment
Censi, P., Randazzo, L.A., Zuddas, P., et al., 2010. Trace element behaviour in seawater of tuffaceous tight oil with separate reservoir and source rock: a case study of tuff
during Etna's pyroclastic activity in 2001: concurrent effects of nutrients and for- reservoir form the Permian Tiaohu Formation in the Santanghu Basin, northwest
mation of alteration minerals. J. Volcanol. Geotherm. Res. 193, 106–116. China. AAPG (Am. Assoc. Pet. Geol.) Bull. 103, 345–367.
Connan, J., 1973. Diagenese naturelle et diagenese artificielle de la matiere organique a Meyer, K.M., Kump, L.R., 2008. Oceanic euxinia in earth history: causes and con-
element vegetaux predominant. In: Tissot, B.P., Bienner, F. (Eds.), Advances in sequences. Annu.rev.earth Palnet Sci 36, 251–288.
Organic Geochemistry. Editions Technip, Paris, pp. 73–95. Meyers, P.A., 1997. Organic geochemical proxies of paleoceanographic, paleolimnologic,
Cranwell, P.A., Eglinton, G., Robinson, N., 1987. Lipids of aquatic organisms as potential and paleoclimatic processes. Org. Geochem. 27, 213–250.
contributors to lacustrine sediments-II. Org. Geochem. 11, 513–527. Meyers, P.A., 2003. Applications of organic geochemistry to paleolimnological re-
Demaison, G.J., Moore, G.T., 1980. Anoxic environments and oil source bed genesis. Org. constructions: a summary of examples from the Laurentian Great Lakes. Org.
Geochem. 2, 9–31. Geochem. 34, 261–289.
Didyk, B.M., Simoneit, B.R.T., Brassell, S.C., et al., 1978. Organic geochemical indicators Morel, F.M.M., Price, N.M., 2003. The biogeochemical cycles of trace metals in the
of palaeoenvironmental conditions of sedimentation. Nature 272, 216–222. oceans. Science 300 (5621), 944.
Ding, X., Liu, G., Zha, M., et al., 2016. Geochemical characterization and depositional Pang, H., Ding, X., Pang, X., et al., 2019. Lower limits of petrophysical parameters al-
environment of source rocks of small fault basin in Erlian Basin, northern China. Mar. lowing tight oil accumulation in the Lucaogou Formation, jimusaer depression,
Petrol. Geol. 69, 231–240. Junggar Basin, western China. Mar. Petrol. Geol. 101, 428–439.
Ding, X., Gao, C., Zha, M., et al., 2017. Depositional environment and factors controlling Peng, Y., Li, Y., Ma, H., et al., 2015. Influencing factors of crude oil properties in
β-carotane accumulation A case study from the jimsar sag, Junggar Basin, north- Lucaogou tight reservoir in jimsar sag, eastern Junggar Basin. Xinjing Pet. Geol. 36,
western China, palaeogeography, palaeoclimatology. Palaeoecology 485, 833–842. 656–659 (in Chinese).
Duggen, S., Croot, P., Schacht, U., et al., 2007. Subduction zone volcanic ash can fertilize Peters, K.E., 1986. Guidelines for evaluating petroleum source rock using programmed
the surface ocean and stimulate phytoplankton growth: evidence from biogeochem- pyrolysis. AAPG Bull. 70, 318–329.
ical experiments and satellite data. Geophys. Res. Lett. 34, 95–119. Peters, K.E., Cassa, M.R., 1994. Applied source rock geochemistry: chapter 5: Part II.
Eglinton, G., Hamilton, R.J., 1967. Leaf epicuticular waxes. Science 156, 1322–1335. Essential elements. In: Magoon, L.B., Dow, W.G. (Eds.), The Petroleum System-From
Eugster, H.P., Surdam, R.C., 1973. Depositional environment of the Green River forma- Source to Trap. American Association of Petroleum Geologists Bulletin, pp. 93–120.
tion: a preliminary report. Bull. Geol. Soc. Am. 86, 319–334. Peters, K.E., Fraser, T.H., Amris, W., et al., 1999. Geochemistry of crude oils from eastern
Grynberg, M.E., Papava, D., Shengelia, M., Takaishvili, A., A, et al., 1993. Petrophysical Indonesia. AAPG (Am. Assoc. Pet. Geol.) Bull. 83, 1927–1942.
characteristics of the middle ecoene laumontite tuff reservoir, Samgori field, Republic Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. second ed. The Biomarker Guide:
of Georgia. J. Pet. Geol. 16, 313–322. Biomarkers and Isotopes in Petroleum Exploration and Earth History, vol. 2
Hao, F., Zhou, X., Zhu, Y., Yang, Y., 2011. Lacustrine source rock deposition in response Cambridge University Press, Cambridge.
to coevolution of environments and organisms controlled by tectonic subsidence and Philp, R.P., Chen, J.H., Fu, J.M., 1992. A geochemical investigation of crude oils and
climate, Bohai Bay basin, China. Org. Geochem. 42, 323–339. source rocks from Biyang Basin, China. Org. Geochem. 18, 933–945.
Harris, N.B., Freeman, K.H., Pancost, R.D., et al., 2004. The character and origin of la- Pickel, W., Kus, J., Flores, D., et al., 2017a. Classification of liptinite-ICCP system 1994.
custrine source rocks in the Lower Cretaceous synrift section, Congo Basin, west Int. J. Coal Geol. 169, 40–61.
Africa. AAPG (Am. Assoc. Pet. Geol.) Bull. 88, 1163–1184. Pickel, W., Kus, J., Flores, D., 2017b. Classification of liptinite - ICCP system 1994. Int. J.
Hedges, J.I., Keil, R.G., 1995. Sedimentary organic matter preservation: an assessment Coal Geol. 169, 40–61.
and speculative synthesis. Mar. Chem. 49, 81–115. Qiu, Z., Shi, Z., Dong, D., et al., 2016. Geological characteristics of source rock and re-
Holba, A.G., Dzou, L.I., Wood, G.D., et al., 2003. Application of tetracyclic polyprenoids servoir of tight oil and its accumulation mechanism: a case study of Permian
as indicators of input from fresh-brackish water environments. Org. Geochem. 34, Lucaogou Formation in Jimusar sag, Junggar Basin. Petrol. Explor. Dev. 43 (6),
441–469. 1013–1024.
Huang, W.Y., Meinschein, W.G., 1979. Sterols as ecological indicators. Geochem. Seifert, W.K., Moldowan, J.M., 1978. Applications of steranes, terpanes and monoaro-
Cosmochim. Acta 43, 739–745. matics to the maturation, migration and source of crude oils. Geochem. Cosmochim.
Huang, D., Li, J., Zhou, X., et al., 1984. Evolution and Hydrocarbon Generation Acta 42, 77–95.
Mechanism of Terrestrial Organic Matter. Petroleum industry press, Beijing, pp. Shao, Y., Yang, Y., Wan, M., et al., 2015. Sedimentary characteristic and facies evolution
25–29 (in Chinese). of permian Lucaogou Formation in jimsar sag, Junggar Basin. Xinjing Pet. Geol. 36,
Jiang, Y., Liu, Y., Yang, Z., et al., 2015. Characteristics and origin of tuff-type tight oil in 635–641 (in Chineses).
jimusar depression, Junggar Basin, NW China. Petrol. Explor. Dev. 42, 741–749. Sinninghe, Damste, J.S., Kenig, F., Koopmans, M.P., et al., 1995. Evidence for gamma-
Jones, M.T., Gislason, S.R., 2008. Rapid releases of metal salts and nutrients following the cerane as an indicator of water column stratification. Geochem. Cosmochim. Acta 59,
deposition of volcanic ash into aqueous environments. Geochem. Cosmochim. Acta 1895–1900.
72 (15), 3661–3680. Sinton, C.W., Duncan, R.A., 1997. Potential links between ocean plateau volcanism and
Katz, B.J., 1990. Controls on distribution of lacustrine source rocks through time and global ocean anoxia at the Cenomanian-Turonian boundary. Econ. Geol. Bull. Soc.
space. In: Katz, B.J. (Ed.), Lacustrine Basin Exploration: Case Studies and Modern Econ. Geol. 92, 836–842.
Analogs. vol. 50. American Association of Petroleum Geologists Memoir, pp. 61–76. Summons, R.E., Hope, J.M., Swart, R., Walter, M.R., 2008. Origin of Nama basin bitumen
Katz, B.,J., 1994. Petroleum source rocks-an introduction overview. In: Katz, B.J. (Ed.), seeps: petroleum derived from a Permian lacustrine source rock traversing south-
Petroleum Source Rock, pp. 1–8. western Gondwana. Org. Geochem. 39, 589–607.
Katz, B.J., 1995. A survey of rift basin source rocks. In: Lambiase, J.J. (Ed.), Hydrocarbon Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence, second ed. Springer
Habitat of Rift Basins. vol. 80. Geological Society (London) Special Publication, pp. Verlag, Berlin.
213–242. Turgeon, S.C., Creaser, R.A., 2008. Cretaceous oceanic anoxic event 2 triggered by a
Katz, B.J., 2005. Controlling factors on source rock development-a review of productivity, massive magmatic episode. Nature 454 (7202), 323–326.
preservation, and sedimentation rate. In: Harris, N.B. (Ed.), The Deposition of Uematsu, M., Toratani, M., Kajino, M., et al., 2004. Enhancement of primary productivity
Organic Carbon Rich Sediments: Models, Mechanisms, and Consequences. Society of in the western North Pacific caused by the eruption of the Miyake-jima Volcano.
Sedimentary Geology, pp. 7–16. Geophys. Res. Lett. 31 (6), 177–182.
Kelts, K., 1988. Environments of deposition of lacustrine petroleum source rocks: an in- Venkatesan, M.I., 1989. Tetrahymanol: its widespread occurrence and geochemical sig-
troduction. In: Fleet, A.J., Kelts, K., Talbot, M.R. (Eds.), Lacustrine Petroleum Source nificance. Geochem. Cosmochim. Acta 53, 3095–3101.
Rocks. vol. 40. Geological Society Special Publication, pp. 3–26. Volkman, J.K., 1983. A review of sterol markers for marine and terrigenous organic-
Keym, M., Dieckmann, V., Horsfield, B., Erdmann, M., Galimberti, R., Kua, L.C., Leith, L., matter. Org. Geochem. 9, 83–99.
Podlaha, O., 2006. Source rock heterogeneity of the upper Jurassic Draupne Volkman, J.K., 2003. Sterols in microorganisms. Appl. Microbiol. Biotechnol. 60,
Formation, north Viking Graben, and its relevance to petroleum generation studies. 495–506.
Org. Geochem. 37, 220–243. Volkman, J.K., Barrett, S.M., Blackburn, S.I., et al., 1998. Microalgal biomarkers: a review
Knoll, A.H., Summons, R.E., Waldbauer, J.R., Zumberge, J.E., 2007. The geological suc- of recent research developments. Org. Geochem. 29, 1163–1179.
cession of primary producers in the oceans. In: Falkowski, P., Knoll, A.H. (Eds.), Wang, P., Pan, J., Wei, D., et al., 2011. A new type of hydrocarbon source rock-sedi-
Evolution of Primary Producers in the Sea. Academic Press, Boston, pp. 133–163. mentary tuff. J. Xi'an Shiyou Univ. 26, 19–22 (in Chinese).
Kuang, L., Tang, Y., Lei, D., Chang, et al., 2012. Formation conditions and exploration Wang, X., Sun, L., Zhu, R., et al., 2015. Application of charging effects in evaluating
potential of tight oil in the Permian saline lacustrine dolomitic rock, Junggar Basin, storage space of tight reservoirs: a case study from Permian Lucaogou Formation in
NW China. Petrol. Explor. Dev. 39, 700–711. Jimusar sag, Junggar Basin, NW China. Petrol. Explor. Dev. 42, 516–524.
Kuang, L., Hu, W., Wang, X., et al., 2013. Research of the Tight Oil Reservoir in the Wei, C., Wei, D., Huang, L., et al., 2013. Geochemical characteristics of Permian heavy oil
Lucaogou Formation in Jimusar Sag: Analysis of Lithology and Porosity and its origin in Jimsar sag, Xinjiang Province. Nat. Gas Geosci. 23, 135–140 (in
Characteristics. vol. 19. pp. 529–535 (in Chinese). Chinese).
Kurenkov, I.I., 1966. The influence of volcanic ash fall on biological processes in a lake. Wu, H., Hu, W., Tang, Y., 2017. The impact of organic fluids on the carbon isotopic
Limnol. Oceanogr. 11, 426–429. compositions of carbonate-rich reservoirs: case study of the Lucaogou Formation in
Langmann, B., Zakšek, K., Hort, M., et al., 2010. Volcanic ash as fertiliser for the surface the Jimusaer Sag, Junggar Basin, NW China. Mar. Petrol. Geol. 85, 136–150.
ocean. Atmos. Chem. Phys. 10, 3891–3899. Xiang, Baoli, Zhang, Ni, Ma, Wanyun, et al., 2015. Multiple-stage migration and accu-
Li, F., Wang, M., Liu, S., et al., 2019. Pore characteristics and influencing factors of dif- mulation of Permian lacustrine mixed oils in the central Junggar Basin (NW China).
ferent types of shales. Mar. Petrol. Geol. 102, 391–401. Mar. Petrol. Geol. 59, 187–201.
Liang, C., Jiang, Z., Cao, Y., et al., 2018. Sedimentary characteristics and origin of la- Yang, Y., Qiu, L., Wan, M., et al., 2018. Depositional model for a salinized lacustrine
custrine organic-rich shales in the salinized Eocene Dongying Depression. 130 (1–2), basin: the permian Lucaogou Formation, jimsar sag, Junggar Basin, NW China. J.

705
X. Ding, et al. Journal of Petroleum Science and Engineering 179 (2019) 696–706

Asian Earth Sci. https://doi.org/10.1016/j.jseaes.2018.08.021. Zhou, Z., Sheng, G., Min, Y., et al., 1989. A primary study of tuffaceous source rock by
Zhang, S., Cao, Y., Liu, K., et al., 2019. Characterization of lacustrine mixed fine-grained organic geochemistry. Acta Sedimentol. Sin. 7, 3–9 (in Chinese).
sedimentary rocks using coupled chemostratigraphic-petrographic analysis: a case Zou, C., Tao, S., Hou, L., et al., 2013. Unconventional Petroleum Geology, second ed.
study from a tight oil reservoir in the Jimusar Sag, Junggar Basin. Mar. Petrol. Geol. Geological Publishing House, Beijing.
99, 453–472.

706

You might also like