Next Article in Journal
Electromagnetic Torque Analysis and Structure Optimization of Interior Permanent Magnet Synchronous Machine with Air-Gap Eccentricity
Previous Article in Journal
Quantitative Characterization of Pore Structure Parameters in Coal Based on Image Processing and SEM Technology
Previous Article in Special Issue
Study on the Tight Gas Accumulation Process and Model in the Transition Zone at the Margin of the Basin: A Case Study on the Permian Lower Shihezi Formation, Duguijiahan Block, Ordos Basin, Northern China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Importance of Laminae for China Lacustrine Shale Oil Enrichment: A Review

1
School of Geosciences, China University of Petroleum, Qingdao 266580, China
2
Key Laboratory of Tectonics and Petroleum Resources, Ministry of Education, China University of Geosciences, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(4), 1661; https://doi.org/10.3390/en16041661
Submission received: 20 December 2022 / Revised: 14 January 2023 / Accepted: 1 February 2023 / Published: 7 February 2023

Abstract

:
The laminar structure of shale system has an important influence on the evaluation of hydrocarbon source rock quality, reservoir quality, and engineering quality, and it is receiving increasing attention. A systematic study of the lamina structure is not only of great scientific significance but also of vital practical importance for shale oil production. In this paper, the identification and description classification of shale laminae are first reviewed. Multiple scales and types indicate that a combination of different probe techniques is the basis for an accurate evaluation of shale laminar characteristics. The influence of laminae on shale reservoir, oil-bearing, mobility, and fracability properties is discussed systematically. A comparative analysis shows that shale systems with well-developed lamination facilitate the development of bedding fractures, thus improving the shale storage space. The average pore size and pore connectivity are also enhanced. These factors synergistically control the superior retention and flow capacity of shale oil in laminated shales. In such conditions, the high production of shale oil wells can still be achieved even if complex networks of fracturing cracks are difficult to form in shale systems with well-developed lamination. This work is helpful to reveal the enrichment mechanism of shale oil and clarify the high-yield law of hydrocarbons, so as to guide the selection of sweet spots.

1. Introduction

Compared to conventional petroleum systems, shale strata exhibit the characteristics of extremely low porosity and ultralow permeability. However, the development of horizontal drilling and multistage hydraulic fracturing technologies has allowed the commercial extraction of hydrocarbons from tight shale deposits that were unprofitable decades ago [1,2,3,4,5,6]. In North America, drilling in the Bakken shale and the Eagle Ford shale confirmed the availability of hydrocarbon liquid resources [1]. The shale oil production in the United States reached an annual level of 2.349 billion barrels in 2018, accounting for 64.7% of the total oil production. This value is expected to reach 9.46 million barrels per day in 2040, accounting for 67.3% of the total US oil production [7,8,9,10]. Following in the footsteps of the US, China is also aggressively examining its unconventional resource potential targeting shale oil plays [1,8,9]. Large-scale shale oil exploration in diverse petroliferous basins, e.g., Junggar Basin, Ordos Basin, Jianghan Basin, Bohai Bay Basin, and Songliao Basin, has revealed that China contains abundant shale oil resources [7,8,9,10,11,12]. The estimated total mass of the technically recoverable resources of shale oil in China is between 7.4 × 109 and 3.7 × 1010 t [7,12]. Currently, an annual shale oil production of about 160 × 104 t has been realized, which makes it possible for shale oil to provide new opportunities for strengthening the energy security of China [8,10,12].
Compared to the North America, the petroleum in China is mostly extracted from lacustrine shale instead of marine shale [8,9,13]. It is worth noting that lake basin sedimentary phases are diverse, with small sedimentation scales and shallow water depth, and they are more influenced by paleoclimatic changes [1,14,15]. Hence, it is easy to form a variety of laminar structures in lacustrine shale systems [16,17,18,19,20,21,22]. More importantly, the expansion of works has realized that laminated shales usually have high total organic carbon (TOC) content and superior reservoir quality. Exploration practice from the Shengli Geophysical Research Institute of Sinopec has shown that shale oil produced by conventional measures mainly comes from laminated shales. Statistics on shale oil-producing wells (more than 300 wells) in the Jiyang Depression indicate that laminated shales with industrial oil flows account for approximately 70% of total pay intervals [15,22]. In other words, the laminated structure is also closely related to oil production [11,21,22,23]. With these results, it is urgent to understand the development characteristics of shale laminae and systematically summarize its impact on oil molecule accumulation, so as to realize commercial production of continental shale oil in China.
Starting from the identification of laminae in shale reservoirs, this paper first summarizes the current common division schemes of lamina types. After that, the effects of laminae structure on shale reservoir, oil-bearing, and mobility properties are discussed. Furthermore, the relationship between laminae and shale fracability is preliminarily analyzed. In general terms, this work is helpful to reveal the enrichment and high-yield mechanism of hydrocarbons from tight unconventional reservoirs.

2. Identification of Multiscale Laminae

The lamina is the smallest megascopic layer without internal layers in fine-grained sedimentary rocks [24,25]. The thickness is generally less than 1 cm [11,26,27]. The mineral composition within a single lamina is relatively uniform [27]. Generally, laminae are interpreted to form in response to small-scale fluctuations within a single flow or depositional event in the rates of the controlling processes, e.g., seasonal growth of planktic or benthic organisms [11,15]. In addition, some laminae may be associated with late diagenetic transformation, such as coarse-grained calcite laminae, which is widely observed in the Shahejie Formation shale in the Dongying sag [16,22]. In such cases, the lamina may have multiscale characteristics in different observational settings. Lamellar features ranging from macroscopic (meter scale) to microscopic (mm to μm scale) have been extensively reported by previous studies [11,17,25,28,29,30] (Figure 1). In general, the meter and centimeter lamina structures reflect variations of vertical superposition lithology, while millimeter and micrometer lamina structures reflect mineral superimposition patterns [31,32].
Core observation can directly depict the centimeter- to millimeter-scale laminae, as well as obtain various superposition methods of different mineral compositions on this scale [27,29] (Figure 1B). The submillimeter- to micron-scale lamellar structures can be further quantitatively described by using thin sections and scanning electron microscopy (SEM) (Figure 1C,D). However, core and microscopic data are limited and often discontinuous, making it difficult to effectively predict the vertical distribution of laminae in a whole well [30,33]. Conversely, well-log data, e.g., a series of conventional logs and image logs, offer special advantages for continuous evaluation [25,30,34]. In particular, image logging has been applied to the identification of laminae on the subcentimeter scale because of its high vertical resolution (~5mm) (Figure 1A). For example, Wang et al. [30] and Pang et al. [25] offered the characteristics of shale multiscale laminae of Fengcheng Formation in Mahu Sag and Lucaogou Formation in Jimusar Sag through imaging logging, respectively, which provided insights for the prediction of sweet spots in the study area. It should be noted that the resolution of the image tool is limited (the vertical resolution is usually 5 mm), which can lead to the loss of microscopic lamina information [30,34]. In other words, the laminated structure investigated by imaging logs is strictly more of a laminar combination, which may contain a large number of basic laminae on the μm scale. Overall, the joint use of multiple methods is the key to the effective evaluation of multiscale laminae (Figure 1).

3. Descriptive Classification of Shale Laminae

Currently, a number of categories have been used to divide the shale laminae [27,29,30,35,36,37,38]. Among them, mineral composition is the most common basis for classification. In such cases, with the application of image observation, clay mineral lamination, felsic lamination, carbonate lamination, and organic matter lamination are widely recognized in the lacustrine shale systems of China (Figure 1) [20,21,38,39,40]. Furthermore, the parameters of lamina, e.g., continuity, shape, and geometry, are used to capture lamination (Figure 2). With these results, three major categories and 12 subclasses of laminae are proposed [35,37], and common lamination patterns and their corresponding images are shown in Figure 2. In addition, Shi et al. [22] divided laminae into thin parallel laminae, thick parallel laminae, wavy laminae, lenticular laminae, sandy laminae, and weak laminae according to the laminar shapes and structures. Moreover, Dong et al. [41] determined the thickness of bedding through core observation and thin section identification, classifying laminar spacing >50 cm as massive, 10–50 cm as layered, 1–10 cm as thin laminae, 1 mm–1 cm as laminae, and <1 mm as sheet. The demarcation point between laminae and layers, according to Liu et al. [27], was set at 1 mm (laminae < 1 mm, layer ≥ 1 mm). Similarly, the value of 1 cm to distinguish lamina and layered structure was implemented in the work of Pang et al. [25]. The aforementioned discussion indicates that the laminae have gradually undergone a transformation in terms of classification and characterization from a single feature to a comprehensive feature [17,42,43]. Laminae can no longer be accurately described solely by virtue of their thickness, shape, and mineral composition.

4. The Role of Laminae for Shale Oil Enrichment and High Yield

4.1. Control of Laminae on Reservoir Property

The occurrence states of oil molecules in shale reservoirs including free oil occurs in pores and microfractures, adsorbed oil storage in kerogen and on the mineral particle surface, and a small amount of oil that dissolves in the kerogen [44,45,46,47]. The work of Chen et al. [48] pointed out that about 80% of shale oil is distributed in macropores of shale reservoir. Hence, the pore structure, including pore size, pore volume, and porosity, possesses an important influence on shale oil storage capacity [49,50,51,52,53]. In addition, previous studies have indicated that the contact boundary between various laminae belongs to the mechanically weak surface of shale rock, which easily forms laminar fractures during geological evolution, thus improving the reservoir’s physical properties [8,9,12,54,55,56]. For example, using NMR and helium porosity tests, Wang et al. [57] and Ning et al. [58] systematically compared shales with different structures in Dongying Sag. Their works pointed out that laminar shales have higher porosity and permeability, followed by layer shales, with massive shales having the lowest (Figure 3A–C). Similar conclusions were demonstrated in [38,59,60,61,62,63,64,65,66]. In particular, based on the quantitative detection of high pressure mercury injection, Bao [60] and Zhang et al. [61] further proposed that the porosity of lamellar shale is mainly provided by pores with a size of 10–30 nm, and the contribution rate is about 32.6–55.2%. Meanwhile, pores with a size of more than 100 nm also have a certain contribution. However, the porosity of massive shale is mainly provided by pores smaller than 10 nm, which generally account for 57.96–99.64% of the total effective pores, with an average value of 78.14%. More importantly, no pores with a size larger than 100 nm could be observed in the massive samples [60,61]. It is worth noting that many studies have also found pores with a diameter of >100 nm in massive shale, but the number is obviously lower than that of laminated shale [57,59,62,67,68,69,70,71].
According to the quantitative statistical results of SEM images, on the other hand, Liu et al. [68] and Zhang [72] identified that various pore types are developed in shales with different structures in the Shahejie Formation of Jiyang Depression (Figure 4). However, the development degree of these pore types is quite different. Among them, lamellar shale is dominated by micron and ultra-micron storage spaces, while massive shale is mainly composed of interlamellar pores and shrinkage pores on the mesopore scale. Pore connectivity is another important parameter to describe the shale reservoir properties [73,74]. Jiang et al. [71] compared the nuclear magnetic resonance and mercury intrusion data of shale, and they concluded that the pore connectivity is ordered as follows: laminated shale > layered shale > massive shale. The same phenomenon was presented using fluid tracer migration [63] and CT scanning [23].
Collectively, the development of laminae promotes the formation of laminar microfractures. In such cases, the pore size of the laminated shale is obviously improved. Moreover, the microfractures can connect the pores, further increasing the porosity and permeability of shale reservoirs [69,71,72]. It is worth noting that the types of laminae are varied (Figure 1). Laminae composed of different minerals also have significantly different characteristics of compaction resistance, which leads to different pore structures between laminae. For example, through quantitative statistics of SEM images of different laminae in the Kongdian Formation shale, Li et al. [24] comprehensively pointed out that the internal pore diameter of carbonate laminae is small due to mineral cementation. At the same time, compared to mixed laminae (mainly composed of clay minerals and siliceous minerals), the work of Xin et al. [21] demonstrated that siliceous laminae possessed superior average porosity and pore size. In addition, Wu et al. [11] found that the porosity and pore size were smaller because of the toughness of the clay laminae and organic matter laminae. As mentioned above, the development characteristics of laminae control the storage conditions of shale reservoirs, which is the basis for shale oil enrichment [47,49,59].

4.2. Influence of Laminae on Oil-Bearing Property

The oil-bearing property of shale refers to the hydrocarbon content retained in reservoirs [46,75]. Likewise, the retention behavior of oil molecules is closely related to the pore structure [46,76,77]. To sum up, from the above discussion, the increase in lamina development seems to be beneficial for the increase in shale oil content [11,53,58]. Ning et al. [58] analyzed Shahejie Formation shales in the Dongying Sag and reported more frequent laminae and a higher oil saturation in the shale. Specifically, the oil saturation of laminated shale is generally above 30%, while that of massive shale is less than 20% (Figure 3D–F). In particular, Shi et al. [22] further compared the difference in oil content in different lamina including thin parallel laminae, thick parallel laminae, wavy laminae, lenticular laminae, weak laminae, and sandy laminae. The first three types of laminae had good continuity and exhibited higher oil saturation values than the latter three types, which had poor laminar continuity. In addition, the free hydrocarbon content (S1) based on Rock-Eval pyrolysis was also used to investigate the oil content in shales [11,20,57,78,79]. The evaluation results also revealed that the oil-bearing property was higher in laminated shale than that in massive shale [21,53,57]. In addition, the fluorescence characteristics are also important parameters to describe the occurrence and enrichment characteristics of shale oil [80,81]. Collectively, the observations of fluorescence thin sections under planepolarized light and fluorescence light show that the fluorescence in laminar shale samples is strong, and it is distributed linearly along the laminated plane or locally concentrated. However, the fluorescence in massive shale is dim and scattered in speckled features [75,82,83,84]. This phenomenon further proves that laminated shale has good oil-bearing properties.
In reality, the laminae can also affect the shale oil content by controlling the expulsion and retention of petroleum [70,82,84]. Generally, the contact surface between different types of laminae is the weak zone in the rock with the lowest rupture strength [58,69]. During the process of hydrocarbon generation and pressurization, these contact surfaces form a large number of microfractures, which promote oil discharge along the microfractures [13,29,47]. In particular, hydrocarbon generation simulation experiments have revealed that under the same or similar conditions, laminated shale has greater oil generation capacity (Figure 5) [85,86,87]. In such cases, the development degree of microfractures and the efficiency of hydrocarbon expulsion in lamellar shales can be further improved [47,57,82]. However, the well-developed laminated fractures can also significantly improve the pore structure and provide important reservoir space for oil storage in a shale system (Figure 3 and Figure 4). Even though a large amount of shale oil may be expelled, as a result, the considerable storage space and oil generation capacity of laminated shale also ensure the demand for shale oil enrichment [88,89]. For this reason, the oil-bearing property of laminated shale is still superior under the background of higher oil discharge efficiency (Figure 5). Overall, the laminar structure affects the hydrocarbon generation and expulsion process of shale oil, thus controlling the oil-bearing characteristics.

4.3. Effect of Laminae on Shale Oil Mobility

The mobility of petroleum in shales is the key to the high yield of shale oil wells [44,58,90,91,92]. On the one hand, the mobility of shale oil depends on the properties of hydrocarbon fluids [47,93,94,95,96]. Complex physical and chemical reactions occur when different complex hydrocarbon compounds flow in micro- to nanopore throat media with different mineral properties [8,97]. In addition to the scale effect, this also includes the adsorption of crude oil components (e.g., gaseous hydrocarbons, saturated hydrocarbons, aromatic hydrocarbons, and asphaltenes) on different minerals [98,99,100]. Generally, adsorbed oil occurs in the form of high-density solidlike or embedded kerogen on the surface of minerals, and it is immovable without additional forces [47]. Considering the properties of oil in shale reservoirs, more light components are beneficial to an increase in movable oil contents [101,102].
On the other hand, reservoir conditions also play specific roles in controlling the flowability of shale oil, and fluid mobility is treated as weaker in small-scale pores (Figure 6) [71,103,104]. For example, the work of Zou et al. [49] found that the lower threshold for oil storage and flow is a pore throat of 20 nm through simulations of a nanoporous template with a controllable pore diameter. At the same time, similar critical pore sizes, e.g., 10 nm [58,90,105], 12.1 nm [96], 30 nm [53], and 40 nm [106], for shale oil flows have been widely reported in related studies based on the SEM observations and quantitative calculation. A molecular dynamics simulation conducted by Wang et al. [107], through adsorption simulations of alkanes onto a graphene surface (oil-wet), proposed that alkanes only flowed after exceeding the critical pore size. The proportion of movable oil gradually increases with increasing pore size. Moreover, this work further realized that the fluidity of oil molecules in slit pores is better than that in circular pores (Figure 6). On the other hand, Cui and Cheng [108] suggested that movable shale oil was also related to the porosity, and the movable oil content increased with increasing porosity [60,89]. As mentioned above, the development of laminae can effectively improve the pore diameter, pore volume, and porosity of shale reservoir (Figure 3 and Figure 4). Compared to massive shales, the laminated shales, thus, possess a higher movable oil proportion (Figure 7). Moreover, Ning et al. [58] and Jiang et al. [71] noted that the development of laminae can also increase the connectivity of the shale reservoir, which is conducive to improving the flow capacity of shale oil.

4.4. Impact of Laminae on Shale Fracability

Compared to conventional reservoirs, the physical characteristics and flow capacity of shale oil reservoirs are quite poor [109,110]. Therefore, using fracturing techniques is essential to connecting matrix pores and achieving commercial productivity of shale oil [111,112]. For this reason, fracability is an important factor to take into account when evaluating shale oil reservoirs [56,113,114]. Studies have demonstrated that shale texture clearly controls the compressive strengths, elastic moduli, and tensile strengths of fine-grained sedimentary rocks [56,109,110,114,115,116,117]. As for the fine-grained sedimentary rocks, laminated shale is often formed by alternating brittle laminae (e.g., carbonate lamina and felsic lamina) and plastic laminae (e.g., clay lamina and organic matter lamina) [11,16,21,22,97]. This situation enhanced the heterogeneity of the rock’s physical properties and tended to concentrate stress [110]. In terms of the impact of structural weak surfaces, interfaces of brittle laminae and plastic laminae are essentially naturally occurring weak planes and frequently the focus of crack propagation [58,69]. An increase in their quantity facilitated the cracking of crack propagation along bedding surfaces [109,115,116], and complex networks of cracks are difficult to form in the process of fracturing (Figure 8A,B) [56,114,118]. Due to their great plasticity in particular, fractures that result from the fracturing of plastic laminae or structurally weak surface can be quickly repaired [56,110,114]. Wang et al. [119] conducted triaxial fracturing modeling experiments and found that lamination limits the height of hydraulic fracturing.
The influence of laminar thickness and laminar continuity on shale fracability was also investigated (Figure 8C–F). Some scholars found that, with the increase in laminar thickness, the weakening degree of laminae to rock decreases, and the brittle index of shale increases [114,118]. Accordingly, a negative correlation between the laminar thickness and shale fracability index (The fracability index is characterized as the ratio of the brittleness index to the fracture toughness. Generally, a higher fracturability index indicates a greater likelihood of the rock fracturing) was recognized (Figure 8C). Meanwhile, a positive correlation with Poisson’s ratio was observed (Figure 8D). Notably, it does not seem that thicker laminae lead to a better fracturing effect. For example, Xu et al. [120] proposed that, when the thickness of shale laminae is moderate (~4 cm), more branching fractures could be recognized, and the fracturing result would be better. As for the laminar continuity, its influence on the fracability of shale is similar to the number of laminae (Figure 8E,F). Actually, this phenomenon is acceptable, as, when the continuity of laminae is great, the fracturing crack could mainly extend along the interface between different lamina, which would hinder the vertical bifurcation and turning of cracks [56,114,116,118].
To sum up, from the above discussion, shale fracturing is reduced with well-developed and strongly continued laminae, but enhanced with laminar thickness (Figure 8). However, it is worth clarifying that laminated shales have high porosity, large pore diameter, good pore connectivity, superior oil content, and great mobility [8,9,13,58,59]. More importantly, interfaces of brittle and plastic laminae are usually the main storage space for shale oil [75,83,84]. The above two factors make it possible to obtain high production even if only simple fractures are formed during fracturing in laminated shales [58,59]. Therefore, laminated shales, e.g., laminated argillaceous shales in the Jiyang Depression and laminated felsic shales in the Huanghua Depression, are usually characterized as the most favorable lithofacies among the continental shales in China [9,58,59,81,121].

5. Conclusions

(1)
Sedimentary lamination is the most important visible sedimentary structure in fine-grained shales. The scales and types of laminae are also multiple. The multiscale laminar structure ranges from macroscopic lithology changes to microcosmic mineral superimposition and can be identified from shale systems. The combination of various probe techniques provides a method for the fine evaluation of multiscale laminae.
(2)
The development of laminae facilitates the formation of microfractures in the shale along the laminar direction. In particular, laminated shales have been identified as generally possessing a greater TOC content and, therefore, superior hydrocarbon generation capacity. These two factors allow the laminated shale to retain and store considerable petroleum even with a higher hydrocarbon expulsion efficiency. In addition, the large pore diameter and good pore connectivity increase the proportion of movable oil. Collectively, the laminated texture controls the shale oil enrichment characteristics.
(3)
The laminar structure also plays an important role in controlling shale fracability. Specifically, fracability declines with increasing laminar number and continuity, but increases with the improvement of laminar thickness. Notably, the high oil content and mobility in the laminated shale allow for high production even if only simple fractures are formed during the fracturing process.
(4)
The laminar structure impacts scale exploration and efficient development of oil. Thus, it is important to strengthen the research on the genetic mechanism and development distribution of laminar for the prediction of shale oil sweet spots. Furthermore, the recognition methods of seismic and logging for the identification and prediction of shale laminae should receive more attention in future work.

Author Contributions

S.X.: conceptualization, validation, resources, data curation, writing—review and editing, supervision, project administration, funding acquisition; Q.G.: methodology, software, formal analysis, investigation, resources, writing—original draft preparation, visualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Shandong Provincial Key Research and Development Program (2020ZLYS08), the National Natural Science Foundation of China (42122017 and 41821002), the independent innovation research program of China University of Petroleum (East China) (21CX06001A), the Open Fund of Key Laboratory of Tectonics and Petroleum Resources (China University of Geosciences), Ministry of Education, China (TPR-2022-19), and the Fundamental Research Funds for National Universities, China University of Geosciences (Wuhan).

Data Availability Statement

The data presented in this study is available on request from the corresponding author.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Katz, B.; Lin, F. Lacustrine basin unconventional resource plays: Key differences. Mar. Pet. Geol. 2014, 56, 255–265. [Google Scholar] [CrossRef]
  2. Li, H. Research progress on evaluation methods and factors influencing shale brittleness: A review. Energy Rep. 2022, 8, 4344–4358. [Google Scholar] [CrossRef]
  3. Li, H.; Zhou, J.; Mou, X.; Guo, H.; Wang, X.; An, H.; Mo, Q.; Long, H.; Dang, C.; Wu, J.; et al. Pore structure and fractal characteristics of the marine shale of the Longmaxi Formation in the Changning Area, Southern Sichuan Basin, China. Front. Earth Sci. 2022, 10, 1018274. [Google Scholar] [CrossRef]
  4. Fan, C.; Li, H.; Qin, Q.; He, S.; Zhong, C. Geological conditions and exploration potential of shale gas reservoir in Wufeng and Longmaxi Formation of southeastern Sichuan Basin, China. J. Pet. Sci. Eng. 2020, 191, 107138. [Google Scholar] [CrossRef]
  5. Li, J.; Li, H.; Yang, C.; Wu, Y.; Gao, Z.; Jiang, S. Geological characteristics and controlling factors of deep shale gas enrichment of the Wufeng-Longmaxi Formation in the southern Sichuan Basin, China. Lithosphere 2022, 12, 4737801. [Google Scholar] [CrossRef]
  6. Gou, Q.; Xu, S.; Hao, F.; Lu, Y.; Shu, Z.; Lu, Y.; Wang, Z.; Wang, Y. Evaluation of the exploration prospect and risk of marine gas shale, southern China: A case study of Wufeng-Longmaxi shales in the Jiaoshiba area and Niutitang shales in the Cen’gong area. GSA Bull. 2022, 134, 1585–1602. [Google Scholar] [CrossRef]
  7. Zou, C.; Pan, S.; Horsfield, B.; Yang, Z.; Hao, S.; Liu, E.; Zhang, L. Oil retention and intrasource migration in the organic-rich lacustrine Chang 7 shale of the Upper Triassic Yanchang Formation, Ordos Basin, central China. AAPG Bull. 2019, 103, 2627–2663. [Google Scholar] [CrossRef]
  8. Jin, Z.; Zhu, R.; Liang, X.; Shen, Y. Several issues worthy of attention in current lacustrine shale oil exploration and development. Pet. Explor. Dev. 2021, 48, 1471–1484. [Google Scholar] [CrossRef]
  9. Jin, Z.; Wang, G.; Liu, G.; Gao, B.; Liu, Q.; Wang, H.; Liang, X.; Wang, R. Research progress and key scientific issues of continental shale oil in China. Acta Pet. Sin. 2021, 42, 821–835, (In Chinese with English abstract). [Google Scholar]
  10. Wang, X.; Zhang, G.; Tang, W.; Wang, D.; Wang, K.; Liu, J.; Du, D. A review of commercial development of continental shale oil in China. Energy Geosci. 2022, 3, 282–289. [Google Scholar] [CrossRef]
  11. Wu, S.; Zhu, R.; Luo, Z.; Yang, Z.; Jiang, X.; Lin, M.; Su, L. Laminar structure of typical continental shales and reservoir quality evaluation in central-western basins in China. China Pet. Explor. 2022, 27, 62–72, (In Chinese with English abstract). [Google Scholar]
  12. Li, Y.; Zhao, Q.; Lyu, Q.; Xue, Z.; Cao, X.; Liu, Z. Evaluation technology and practice of continental shale oil development in China. Pet. Explor. Dev. 2022, 49, 1098–1109. [Google Scholar] [CrossRef]
  13. Zhao, W.; Zhu, R.; Liu, W.; Bian, C.; Wang, K. Enrichment conditions and occurrence features of lacustrine mid-high matured shale oil in onshore China. Earth Sci. Front. 2023, 30, 116–127, (In Chinese with English abstract). [Google Scholar] [CrossRef]
  14. Bohacs, K.M.; Carroll, A.R.; Neal, J.E.; Mankiewicz, P.J. Lake-basin type, source potential, and hydrocarbon character: An integrated sequence-stratigraphic-geochemical framework. In Lake Basins through Space and Time; American Association of Petroleum Geologists: Tulsa, OK, USA, 2000; pp. 3–34. [Google Scholar]
  15. Shi, J.; Jin, Z.; Liu, Q.; Fan, T.; Gao, Z. Sunspot cycles recorded in Eocene lacustrine fine-grained sedimentary rocks in the Bohai Bay Basin, eastern China. Glob. Planet. Change 2021, 205, 103614. [Google Scholar] [CrossRef]
  16. Liang, C.; Cao, Y.; Jiang, Z.; Wu, J.; Song, G.; Wang, Y. Shale oil potential of lacustrine black shale in the Eocene Dongying depression: Implications for geochemistry and reservoir characteristics. AAPG Bull. 2017, 101, 1835–1858. [Google Scholar] [CrossRef]
  17. Li, L.; Huang, B.; Li, Y.; Hu, R.; Li, X. Multi-scale modeling of shale laminas and fracture networks in the Yanchang formation, Southern Ordos Basin, China. Eng. Geol. 2018, 243, 231–240. [Google Scholar] [CrossRef]
  18. Wang, M.; Guan, Y.; Li, C.; Liu, Y.; Liu, W. Qualitative description and full-pore-size quantitative evaluation of pores in lacustrine shale reservoir of Shahejie Formation, Jiyang Depression. Oil Gas Geol. 2018, 39, 1107–1119. [Google Scholar]
  19. Zhang, L.; Chen, Z.; Li, Z.; Zhang, S.; Li, J.; Liu, Q.; Zhu, R.; Zhang, J.; Bao, Y. Structural features and genesis of microscopic pores in lacustrine shale in an oil window: A case study of the Dongying depression. AAPG Bull. 2019, 103, 1889–1924. [Google Scholar] [CrossRef]
  20. Xu, Q.; Zhao, X.; Pu, X.; Han, W.; Shi, Z.; Tian, J.; Zhang, B.; Xin, B.; Guo, P. Characteristics and Control Mechanism of Lacustrine Shale Oil Reservoir in the Member 2 of Kongdian Formation in Cangdong Sag, Bohai Bay Basin, China. Front. Earth Sci. 2021, 9, 783042. [Google Scholar] [CrossRef]
  21. Xin, B.; Zhao, X.; Hao, F.; Jin, F.; Pu, X.; Han, W.; Xu, Q.; Guo, P.; Tian, J. Laminae characteristics of lacustrine shales from the Paleogene Kongdian Formation in the Cangdong Sag, Bohai Bay Basin, China: Why do laminated shales have better reservoir physical properties? Int. J. Coal Geol. 2022, 260, 104056. [Google Scholar] [CrossRef]
  22. Shi, J.; Jin, Z.; Liu, Q.; Zhang, T.; Fan, T.; Gao, Z. Laminar characteristics of lacustrine organic-rich shales and their significance for shale reservoir formation: A case study of the Paleogene shales in the Dongying Sag, Bohai Bay Basin, China. J. Asian Earth Sci. 2022, 223, 104976. [Google Scholar] [CrossRef]
  23. Zhao, X.; Zhou, L.; Pu, X.; Jin, F.; Han, W.; Xiao, D.; Chen, S.; Shi, Z.; Zhang, W.; Yang, F. Geological characteristics of shale rock system and shale oil exploration breakthrough in a lacustrine basin: A case study from the Paleogene 1st sub-member of Kong 2 Member in Cangdong Sag, Bohai Bay Basin, China. Pet. Explor. Dev. 2018, 45, 377–388. [Google Scholar] [CrossRef]
  24. Li, M.; Wu, S.; Hu, S.; Zhu, R.; Meng, S.; Yang, J. Lamination Texture and Its Effects on Reservoir and Geochemical Properties of the Palaeogene Kongdian Formation in the Cangdong Sag, Bohai Bay Basin, China. Minerals 2021, 11, 1360. [Google Scholar] [CrossRef]
  25. Pang, X.J.; Wang, G.W.; Kuang, L.C.; Lai, J.; Gao, Y.; Zhao, Y.; Li, H.; Wang, S.; Bao, M.; Liu, S.; et al. Prediction of multiscale laminae structure and reservoir quality in fine-grained sedimentary rocks: The Permian Lucaogou Formation in Jimusar Sag, Junggar Basin. Pet. Sci. 2022, 19, 2549–2571. [Google Scholar] [CrossRef]
  26. Ingram, R.L. Terminology for the thickness of stratification and parting units in sedimentary rocks. GSA Bull. 1954, 65, 937–938. [Google Scholar] [CrossRef]
  27. Liu, H.; Wang, Y.; Yang, Y.; Zhang, S. Sedimentary environment and lithofacies of fine-grained hybrid in Dongying Sag: A case of fine-grained sedimentary system of the Es4. Earth Sci. 2020, 45, 3543–3555. [Google Scholar]
  28. Anderson, R.Y. Seasonal sedimentation: A framework for reconstructing climatic and environmental change. Geological Society, London. Spec. Publ. 1996, 116, 1–15. [Google Scholar] [CrossRef]
  29. Chen, S.; Zhang, S.; Liu, H.; Yan, J. Discussion on mixing of fine-grained sediments in lacustrine deep water. J. Palaeogeogr. 2017, 19, 271–284. [Google Scholar]
  30. Wang, S.; Wang, G.; Huang, L.; Song, L.; Zhang, Y.; Li, D.; Huang, Y. Logging evaluation of lamina structure and reservoir quality in shale oil reservoir of Fengcheng Formation in Mahu Sag, China. Mar. Pet. Geol. 2021, 133, 105299. [Google Scholar] [CrossRef]
  31. Rickman, R.; Mullen, M.J.; Petre, J.E.; Grieser, B.; Kundert, D. A practical use of shale petrophysics for stimulation design optimization: All shale plays are not clones of the Barnett Shale. In Proceedings of the SPE Annual Technical Conference and Exhibition, Denver, CO, USA, 21—24 September 2008; p. SPE-115258-MS. [Google Scholar] [CrossRef]
  32. Aplin, A.C.; Macquaker, J.H.S. Mudstone diversity: Origin and implications for source, seal, and reservoir properties in petroleum systems. AAPG Bull. 2011, 95, 2031–2059. [Google Scholar] [CrossRef]
  33. Ji, W.; Song, Y.; Jiang, Z.; Chen, L.; Li, Z.; Yang, X.; Meng, M. Estimation of marine shale methane adsorption capacity based on experimental investigations of Lower Silurian Longmaxi formation in the Upper Yangtze Platform, south China. Mar. Pet. Geol. 2015, 68, 94–106. [Google Scholar] [CrossRef]
  34. Lai, J.; Wang, G.; Fan, Z.; Chen, J.; Wang, S.; Fan, X. Sedimentary characterization of a braided delta using well logs: The Upper Triassic Xujiahe formation in central Sichuan basin, China. J. Pet. Sci. Eng. 2017, 154, 172–193. [Google Scholar] [CrossRef]
  35. Campbell, C.V. Lamina, laminaset, bed and bedset. Sedimentology 1967, 8, 7–26. [Google Scholar] [CrossRef]
  36. Jiang, Z.; Chen, D.; Qiu, L.; Liang, H.; Ma, J. Source-controlled carbonates in a small Eocene half-graben lake basin (Shulu Sag) in central Hebei Province, North China. Sedimentology 2007, 54, 265–292. [Google Scholar] [CrossRef]
  37. Lazar, O.R.; Bohacs, K.M.; Macquaker, J.H.S.; Schieber, J.; Demko, T.M. Capturing key attributes of fine-grained sedimentary rocks in outcrops, cores, and thin sections: Nomenclature and description guidelines. J. Sediment. Res. 2015, 85, 230–246. [Google Scholar] [CrossRef]
  38. Liu, G.; Huang, Z.; Jiang, Z.; Chen, J.; Chen, C.; Gao, X. The characteristic and reservoir significance of lamina in shale from Yanchang Formation of Ordos Basin. Nat. Gas Geosci. 2015, 26, 408–411, (In Chinese with English abstract). [Google Scholar]
  39. Shi, Z.; Qiu, Z. Main bedding types of marine fine-grained sediments and their significance for oil and gas exploration and development. Acta Sedimentol. Sin. 2021, 39, 181–196, (In Chinese with English abstract). [Google Scholar]
  40. Liang, C.; Wu, J.; Cao, Y.; Liu, K.; Khan, D. Storage space development and hydrocarbon occurrence model controlled by lithofacies in the Eocene Jiyang Sub-basin, East China: Significance for shale oil reservoir formation. J. Pet. Sci. Eng. 2022, 2015, 110631. [Google Scholar] [CrossRef]
  41. Dong, C.; Ma, C.; Lin, C.; Sun, X.; Yuan, M. A method of classification of shale set. J. China Univ. Pet. 2015, 39, 1–7, (In Chinese with English abstract). [Google Scholar]
  42. Hua, G.L.; Wu, S.T.; Qiu, Z.; Jing, Z.; Xu, J.; Guan, M. Lamination texture and its effect on reservoir properties: A case study of Longmaxi Shale, Sichuan Basin. Acta Sedimentol. Sin. 2021, 32, 1–22. [Google Scholar]
  43. Xing, M.; Chen, L.; Chen, X.; Ji, Y.; Wu, P.; Hu, Y.; Wang, G.; Peng, H. Characteristics, genetic mechanism of marine shale laminae and its significance of shale gas accumulation. J. Cent. South Univ. (Sci. Technol.) 2022, 53, 3490–3508, (In Chinese with English abstract). [Google Scholar]
  44. Jarvie, D.M. Shale resource systems for oil and gas: Part 2—Shale-oil resource systems. In Shale Reservoirs—Giant Resources for the 21st Century: AAPG Memoir 97; American Association of Petroleum Geologists: Tulsa, OK, USA, 2012; pp. 89–119. [Google Scholar]
  45. Liu, X.; Zhang, D. A review of phase behavior simulation of hydrocarbons in confined space: Implications for shale oil and shale gas. J. Nat. Gas Sci. Eng. 2019, 68, 102901. [Google Scholar] [CrossRef]
  46. Zhu, C.; Guo, W.; Li, Y.; Gong, H.; Sheng, J.; Dong, M. Effect of occurrence states of fluid and pore structures on shale oil movability. Fuel 2021, 288, 119847. [Google Scholar] [CrossRef]
  47. Feng, Y.; Xiao, X.; Wang, E.; Sun, J.; Gao, P. Oil retention in shales: A review of the mechanism, controls and assessment. Front. Earth Sci. 2021, 9, 720839. [Google Scholar] [CrossRef]
  48. Chen, G.; Lu, S.; Zhang, J.; Pervukhina, M.; Liu, K.; Wang, M.; Han, T.; Tian, S.; Li, J.; Zhang, Y.; et al. A method for determining oil-bearing pore size distribution in shales: A case study from the Damintun Sag, China. J. Pet. Sci. Eng. 2018, 166, 673–678. [Google Scholar] [CrossRef]
  49. Zou, C.; Jin, X.; Zhu, R.; Gong, G.; Sun, L.; Dai, J.; Meng, D.; Wang, X.; Li, J.; Wu, S.; et al. Do shale pore throats have a threshold diameter for oil storage? Sci. Rep. 2015, 5, 1–6. [Google Scholar] [CrossRef]
  50. Wu, S.T.; Zou, C.N.; Zhu, R.K.; Yuan, X.; Yao, J.; Yang, Z.; Liang, S.; Bai, B. Reservoir quality characterization of upper triassic Chang 7 shale in Ordos basin. Earth Sci. 2015, 40, 1810–1823. [Google Scholar]
  51. Johnson, R.C.; Birdwell, J.E.; Mercier, T.J.; Brownfield, M.E. Geology of tight oil and potential tight oil reservoirs in the lower part of the Green River Formation, Uinta, Piceance, and Greater Green River Basins, Utah, Colorado, and Wyoming. US Geol. Surv. 2016. [Google Scholar] [CrossRef]
  52. Schimmelmann, A.; Lange, C.B.; Schieber, J.; Francus, P.; Ojala, A.E.K.; Zolitschka, B. Varves in marine sediments: A review. Earth-Sci. Rev. 2016, 159, 215–246. [Google Scholar] [CrossRef]
  53. Wang, M.; Ma, R.; Li, J.; Lu, S.; Li, C.; Guo, Z.; Li, Z. Occurrence mechanism of lacustrine shale oil in the Paleogene Shahejie Formation of Jiyang depression, Bohai Bay Basin, China. Pet. Explor. Dev. 2019, 46, 833–846. [Google Scholar] [CrossRef]
  54. Ma, C.; Elsworth, D.; Dong, C.; Lin, C.; Luan, G.; Chen, B.; Liu, X.; Muhammad, J.M.; Muhammad, A.Z.; Shen, Z.; et al. Controls of hydrocarbon generation on the development of expulsion fractures in organic-rich shale: Based on the Paleogene Shahejie Formation in the Jiyang Depression, Bohai Bay Basin, East China. Mar. Pet. Geol. 2017, 86, 1406–1416. [Google Scholar] [CrossRef]
  55. Li, H.; Tang, H.; Qin, Q.; Zhou, J.; Qin, Z.; Fan, C.; Su, P.; Wang, Q.; Zhong, C. Characteristics, formation periods and genetic mechanisms of tectonic fractures in the tight gas sandstones reservoir: A case study of Xujiahe Formation in YB area, Sichuan Basin, China. J. Pet. Sci. Eng. 2019, 178, 723–735. [Google Scholar] [CrossRef]
  56. Xiong, Z.; Wang, G.; Cao, Y.; Liang, C.; Li, M.; Shi, X.; Zhang, B.; Li, J.; Fu, Y. Controlling effect of texture on fracability in lacustrine fine-grained sedimentary rocks. Mar. Pet. Geol. 2019, 101, 195–210. [Google Scholar] [CrossRef]
  57. Wang, Y.; Wang, X.; Song, G.; Liu, H.; Zhu, D.; Zhu, D.; Ding, J.; Yang, W.; Yin, Y.; Zhang, S.; et al. Genetic connection between mud shale lithofacies and shale oil enrichment in Jiyang Depression, Bohai Bay Basin. Pet. Explor. Dev. 2016, 43, 759–768. [Google Scholar] [CrossRef]
  58. Ning, F.; Wang, X.; Hao, X.; Yang, W.; Yin, Y.; Ding, J.; Zhu, D.; Zhu, D.; Zhu, J. Occurrence mechanism of shale oil with different lithofacies in Jiyang Depression. Acta Pet. Sin. 2017, 38, 185–195. [Google Scholar]
  59. Sun, H.Q. Exploration practice and cognitions of shale oil in Jiyang depression. China Pet. Explor. 2017, 22, 1–14. [Google Scholar]
  60. Bao, Y.S. Effective reservoir spaces of Paleogene shale oil in the Dongying Depression, Bohai Bay Basin. Pet. Geol. Expepiment 2018, 40, 480–484, (In Chinese with English abstract). [Google Scholar]
  61. Zhang, L.; Bao, Y.S. Xi, C. Pore Structure Characteristics and Pore Connectivity of Paleogene Shales in Dongying Depression. Xinjiang Pet. Geol. 2018, 39, 134–139, (In Chinese with English abstract). [Google Scholar]
  62. Hu, Q.; Zhang, X.; Meng, X.; Li, Z.; Xie, Z.; Li, M. Characterization of micro-nano pore networks in shale oil reservoirs of Paleogene Shahejie Formation in Dongying Sag of Bohai Bay Basin, East China. Pet. Explor. Dev. 2017, 44, 720–730. [Google Scholar] [CrossRef]
  63. Hu, Q.; Liu, H.; Li, M.; Li, Z.; Yang, R.; Zhang, Y.; Sun, M. Wettability, pore connectivity and fluid-tracer migration in shale oil reservoirs of Paleogene Shahejie Formation in Dongying sag of Bohai Bay Basin, East China. Acta Pet. Sin. 2018, 39, 278–289. [Google Scholar]
  64. Teng, J.; Liu, H.; Qiu, L.; Zhang, S.; Hao, Y.; Tian, F.; Zhu, L.; Fang, Z. Control law of material components of shale oil reservoir on oil-bearing characteristics in Jiyang Depression. Editor. Dep. Pet. Geol. Recovery Effic. 2019, 26, 80–87, (In Chinese with English abstract). [Google Scholar]
  65. Song, M.S. Practice and current status of shale oil exploration in Jiyang Depression. Editorial Dep. Pet. Geol. Recovery Effic. 2019, 26, 1–12, (In Chinese with English abstract). [Google Scholar]
  66. Li, T.; Jiang, Z.; Su, P.; Zhang, X.; Chen, W.; Wang, X.; Ning, C.; Wang, Z.; Xue, Z. Effect of laminae development on pore structure in the lower third member of the Shahejie Shale, Zhanhua Sag, Eastern China. Interpretation 2020, 8, T103–T114. [Google Scholar] [CrossRef]
  67. Chen, S.; Zhang, S.; Wang, Y.; Tan, M. Lithofacies types and reservoirs of Paleogene fine-grained sedimentary rocks Dongying Sag, Bohai Bay Basin, China. Pet. Explor. Dev. 2016, 43, 218–229. [Google Scholar] [CrossRef]
  68. Liu, H.M.; Zhang, S.; Bao, S.Y.; Fang, Z.; Yao, S.; Wang, Y. Geological characteristics and effectiveness of the shale oil reservoir in Dongying sag. Oil Gas Geol. 2019, 40, 512–523. [Google Scholar]
  69. Song, M.; Liu, H.; Wang, Y.; Liu, Y. Enrichment rules and exploration practices of Paleogene shale oil in Jiyang Depression, Bohai Bay Basin, China. Pet. Explor. Dev. 2020, 47, 225–235. [Google Scholar] [CrossRef]
  70. Hu, S.; Zhao, W.; Hou, L.; Yan, Z.; Zhu, R.; Wu, S.; Bai, B.; Jin, X. Development potential and technical strategy of continental shale oil in China. Pet. Explor. Dev. 2020, 47, 877–887. [Google Scholar] [CrossRef]
  71. Jiang, Z.; Li, T.; Gong, H.; Jiang, T.; Chang, J.; Ning, C.; Su, S.; Chen, W. Characteristics of low-mature shale reservoirs in Zhanhua Sag and their influence on the mobility of shale oil. Acta Pet. Sin. 2020, 41, 1587–1600. [Google Scholar]
  72. Zhang, S. Shale oil enrichment elements and geological dessert types in Jiyang Depression. Sci. Technol. Eng. 2021, 21, 504–511, (In Chinese with English abstract). [Google Scholar]
  73. Gou, Q.; Xu, S.; Hao, F.; Yang, F.; Zhang, B.; Shu, Z.; Zhang, A.; Wang, Y.; Cheng, X.; Qing, J.; et al. Full-scale pores and micro-fractures characterization using FE-SEM, gas adsorption, nano-CT and micro-CT: A case study of the Silurian Longmaxi Formation shale in the Fuling area, Sichuan Basin, China. Fuel 2019, 253, 167–179. [Google Scholar] [CrossRef]
  74. Gou, Q.; Xu, S.; Hao, F.; Yang, F.; Shu, Z.; Liu, R. The effect of tectonic deformation and preservation condition on the shale pore structure using adsorption-based textural quantification and 3D image observation. Energy 2021, 219, 119579. [Google Scholar] [CrossRef]
  75. Guan, M.; Liu, X.; Jin, Z.; Lai, J.; Liu, J.; Sun, B.; Liu, T.; Hua, Z.; Xu, W.; Shu, H.; et al. Quantitative characterization of various oil contents and spatial distribution in lacustrine shales: Insight from petroleum compositional characteristics derived from programed pyrolysis. Mar. Pet. Geol. 2022, 138, 105522. [Google Scholar] [CrossRef]
  76. Liu, X.; Lai, J.; Fan, X.; Shu, H.; Wang, G.; Ma, X.; Liu, M.; Guan, M.; Luo, Y. Insights in the pore structure, fluid mobility and oiliness in oil shales of Paleogene Funing Formation in Subei Basin, China. Mar. Pet. Geol. 2020, 114, 104228. [Google Scholar] [CrossRef]
  77. Xu, Y.; Lun, Z.M.; Pan, Z.J.; Wang, H.; Zhou, X.; Zhao, C.; Zhang, D. Occurrence space and state of shale oil: A review. J. Pet. Sci. Eng. 2022, 211, 110183. [Google Scholar] [CrossRef]
  78. Li, Z.; Qian, M.; Li, M.; Jiang, Q.; Liu, P.; Rui, X.; Cao, T.; Pan, Y. Oil content and occurrence in low-medium mature organic-rich lacustrine shales: A case from the 1st member of the Eocene-Oligocene Shahejie Formation in Well Luo-63 and Yi-21, Bonan Subsag, Bohai Bay Basin. Oil Gas Geol. 2017, 38, 448–456. [Google Scholar]
  79. Zeng, X.; Cai, J.; Dong, Z.; Wang, X.; Hao, Y. Sedimentary characteristics and hydrocarbon generation potential of mudstone and shale: A case study of Middle Submember of Member 3 and Upper Submember of Member 4 in Shahejie Formation in Dongying sag. Acta Pet. Sin 2017, 38, 31–43. [Google Scholar]
  80. Nikolaev, M.Y.; Kazak, A.V. Liquid saturation evaluation in organic-rich unconventional reservoirs: A comprehensive review. Earth Sci. Rev. 2019, 194, 327–349. [Google Scholar] [CrossRef]
  81. Zhao, X.; Pu, X.; Zhou, L.; Jin, F.; Han, G.; Shi, Z.; Han, W.; Ding, Y.; Zhang, W.; Wang, G.; et al. Enrichment theory, exploration technology and prospects of shale oil in lacustrine facies zone of deep basin: A case study of the Paleogene in Huanghua Depression, Bohai Bay Basin. Acta Pet. Sin. 2021, 42, 143–162. [Google Scholar]
  82. Du, J.; Hu, S.; Pang, Z.; Lin, S.; Hou, L.; Zhu, R. The types, potentials and prospects of continental shale oil in China. China Pet. Explor. 2019, 24, 560–568. [Google Scholar]
  83. Zhou, L.; Zhao, X.; Chai, G.; Jiang, W.; Pu, X.; Wang, X.; Han, W.; Guan, Q.; Feng, J.; Liu, X. Key exploration & development technologies and engineering practice of continental shale oil: A case study of Member 2 of Paleogene Kongdian Formation in Cangdong Sag, Bohai Bay Basin, East China. Pet. Explor. Dev. 2020, 47, 1138–1146. [Google Scholar]
  84. Zhao, W.; Zhu, R.; Hu, S.; Hou, L.; Wu, S. Accumulation contribution differences between lacustrine organic-rich shales and mudstones and their significance in shale oil evaluation. Pet. Explor. Dev. 2020, 47, 1160–1171. [Google Scholar] [CrossRef]
  85. Ma, L.; Cheng, K.; Liu, D.; Xiong, Y. Laminar algal distribution characteristics of lower Cretaceous and the relation to oil-gas of Jiuquan Basin. Acta Sedimentol. Sin. 2007, 25, 147–153, (In Chinese with English abstract). [Google Scholar]
  86. Zhang, S.C.; Zhang, L.Y.; Zha, M. Research on simulation of hydrocarbon expulsion difference in lacustrine source rocks: A case study of Paleogene Es3 member in the Dongying Depression. Pet. Geol. Recovery Effic. 2009, 16, 32–35, (In Chinese with English abstract). [Google Scholar]
  87. Jin, Q.; Zhang, H.J.; Cheng, F.Q.; Xu, J. Stimulations on generation, expulsion and retention of liquid hydrocarbons in source rocks deposited in lacustrine basin and their significance in petroleum geology. J. China Univ. Pet. (Ed. Nat. Sci.) 2019, 43, 44–52, (In Chinese with English abstract). [Google Scholar]
  88. Xi, K.; Li, K.; Cao, Y.; Lin, M.; Niu, X.; Zhu, R.; Wei, X.; You, Y.; Liang, X.; Feng, S. Laminae combination and shale oil enrichment patterns of Chang 73 sub-member organic-rich shales in the Triassic Yanchang Formation, Ordos Basin, NW China. Pet. Explor. Dev. 2020, 47, 1342–1353. [Google Scholar] [CrossRef]
  89. Han, W.; Zhao, X.; Jin, F.; Pu, X.; Chen, S.; Mou, L.; Zhang, W.; Shi, Z.; Wang, H. Sweet spots evaluation and exploration of lacustrine shale oil of the 2nd member of Paleogene Kongdian Formation in Cangdong sag, Dagang Oilfield, China. Pet. Explor. Dev. 2021, 48, 1–10. [Google Scholar] [CrossRef]
  90. Li, M.; Chen, Z.; Ma, X.; Cao, T.; Qian, M.; Jiang, Q.; Tao, G.; Li, Z.; Song, G. Shale oil resource potential and oil mobility characteristics of the eocene-oligocene Shahejie Formation, Jiyang super-depression, Bohai Bay Basin of China. Int. J. Coal Geol. 2019, 204, 130–143. [Google Scholar] [CrossRef]
  91. Zhang, P.; Lu, S.; Lin, Z.; Duan, H.; Chang, X.; Qiu, Y.; Fu, Q.; Zhi, Q.; Wang, J.; Huang, H. Key Oil Content Parameter Correction of Shale Oil Resources: A Case Study of the Paleogene Funing Formation, Subei Basin, China. Energy Fuels 2022, 36, 5316–5326. [Google Scholar] [CrossRef]
  92. Wang, E.; Li, C.; Feng, Y.; Song, Y.; Guo, T.; Li, M.; Chen, Z. Novel method for determining the oil moveable threshold and an innovative model for evaluating the oil content in shales. Energy 2022, 239, 121848. [Google Scholar] [CrossRef]
  93. Teklu, T.W.; Alharthy, N.; Kazemi, H.; Yin, X.; Graves, R.M. Phase behavior and minimum miscibility pressure in nanopores. SPE Reserv. Eval. Eng. 2014, 17, 396–403. [Google Scholar] [CrossRef]
  94. Hu, T.; Pang, X.; Jiang, S.; Wang, Q.; Zheng, X.; Ding, X.; Zhao, Y.; Zhu, C.; Li, H. Oil content evaluation of lacustrine organic-rich shale with strong heterogeneity: A case study of the Middle Permian Lucaogou Formation in Jimusaer Sag, Junggar Basin, NW China. Fuel 2018, 221, 196–205. [Google Scholar] [CrossRef]
  95. Sobecki, N.; Nieto-Draghi, C.; Di Lella, A.; Ding, D.Y. Phase behavior of hydrocarbons in nano-pores. Fluid Phase Equilibria 2019, 497, 104–121. [Google Scholar] [CrossRef]
  96. Zhu, X.; Cai, J.; Liu, Q.; Li, Z.; Zhang, X. Thresholds of petroleum content and pore diameter for petroleum mobility in shale. AAPG Bull. 2019, 103, 605–617. [Google Scholar] [CrossRef]
  97. Liang, C.; Cao, Y.; Liu, K.; Jiang, Z.; Wu, J.; Hao, F. Diagenetic variation at the lamina scale in lacustrine organic-rich shales: Implications for hydrocarbon migration and accumulation. Geochim. Cosmochim. Acta 2018, 229, 112–128. [Google Scholar] [CrossRef]
  98. Ribeiro, R.C.; Correia, J.C.G.; Seidl, P.R. The influence of different minerals on the mechanical resistance of asphalt mixtures. J. Pet. Sci. Eng. 2009, 65, 171–174. [Google Scholar] [CrossRef]
  99. Mohammadi, M.; Sedighi, M. Modification of Langmuir isotherm for the adsorption of asphaltene or resin onto calcite mineral surface: Comparison of linear and non-linear methods. Prot. Met. Phys. Chem. Surf. 2013, 49, 460–470. [Google Scholar] [CrossRef]
  100. Tian, S.; Erastova, V.; Lu, S.; Greenwell, H.C.; Underwood, T.R.; Xue, H.; Zeng, F.; Chen, G.; Wu, C.; Zhao, R. Understanding model crude oil component interactions on kaolinite silicate and aluminol surfaces: Toward improved understanding of shale oil recovery. Energy Fuels 2018, 32, 1155–1165. [Google Scholar] [CrossRef]
  101. Jarvie, D.M. Components and processes affecting producibility and commerciality of shale resource systems. Geol. Acta 2014, 12, 307–325. [Google Scholar]
  102. Fan, B.; Shi, L. Deep-lacustrine shale heterogeneity and its impact on hydrocarbon generation, expulsion, and retention: A case study from the upper triassic Yanchang Formation, Ordos Basin, China. Nat. Resour. Res. 2019, 28, 241–257. [Google Scholar] [CrossRef]
  103. Dillinger, A.; Esteban, L. Experimental evaluation of reservoir quality in Mesozoic formations of the Perth Basin (Western Australia) by using a laboratory low field nuclear magnetic resonance. Mar. Pet. Geol. 2014, 57, 455–469. [Google Scholar] [CrossRef]
  104. Ning, C.; Ma, Z.; Jiang, Z.; Su, S.; Li, T.; Zheng, L.; Wang, G.; Li, F. Effect of shale reservoir characteristics on shale oil movability in the lower third member of the Shahejie Formation, Zhanhua Sag. Acta Geol. Sin. Engl. Ed. 2020, 94, 352–363. [Google Scholar] [CrossRef]
  105. Sun, C.; Yao, S.; Li, J.; Liu, B.; Liu, H.; Xie, Z. Characteristics of pore structure and effectiveness of shale oil reservoir space in Dongying Sag, Jiyang Depression, Bohai Bay Basin. J. Nanosci. Nanotechnol. 2017, 17, 6781–6790. [Google Scholar] [CrossRef]
  106. Pan, Y.; Huang, Z.; Guo, X.; Liu, B.; Wang, G.; Xu, X. Study on the pore structure, fluid mobility, and oiliness of the lacustrine organic-rich shale affected by volcanic ash from the Permian Lucaogou Formation in the Santanghu Basin, Northwest China. J. Pet. Sci. Eng. 2022, 208, 109351. [Google Scholar] [CrossRef]
  107. Wang, S.; Feng, Q.; Javadpour, F.; Xia, T.; Li, Z. Oil adsorption in shale nanopores and its effect on recoverable oil-in-place. Int. J. Coal Geol. 2015, 147, 9–24. [Google Scholar] [CrossRef]
  108. Cui, J.; Cheng, L. A theoretical study of the occurrence state of shale oil based on the pore sizes of mixed Gaussian distribution. Fuel 2017, 206, 564–571. [Google Scholar] [CrossRef]
  109. Tan, P.; Jin, Y.; Han, K.; Hou, B.; Chen, M.; Guo, X.; Gao, J. Analysis of hydraulic fracture initiation and vertical propagation behavior in laminated shale formation. Fuel 2017, 206, 482–493. [Google Scholar] [CrossRef]
  110. Ma, C.; Dong, C.; Lin, C.; Elsworth, D.; Luan, G.; Sun, X.; Liu, X. Influencing factors and fracability of lacustrine shale oil reservoirs. Mar. Pet. Geol. 2019, 110, 463–471. [Google Scholar] [CrossRef]
  111. Sheng, Q.; Li, W. Evaluation method of shale fracability and its application in Jiaoshiba area. Prog. Geophys. 2016, 31, 1473–1479. [Google Scholar]
  112. Osiptsov, A.A. Fluid mechanics of hydraulic fracturing: A review. J. Pet. Sci. Eng. 2017, 156, 513–535. [Google Scholar] [CrossRef]
  113. Chong, K.K.; Grieser, W.V.; Passman, A.; Tamayo, C.H.; Modeland, N.; Burke, B. A completions guide book to shale-play development: A review of successful approaches towards shale-play stimulation in the last two decades. In Proceedings of the Canadian Unconventional Resources and International Petroleum Conference, Calgary, AB, Canada, 19–21 October 2010; pp. SPE-133874-MS. [Google Scholar] [CrossRef]
  114. Xiong, Z.; Cao, Y.; Wang, G.; Liang, C.; Shi, X.; Li, M.; Fu, Y.; Zhao, S. Influence of laminar structure differences on the fracability of lacustrine fine-grained sedimentary rocks. Acta Pet. Sin. 2019, 40, 74–85. [Google Scholar]
  115. Zhao, W.; Hou, G.; Zhang, J.; Feng, S.; Ju, W.; You, Y.; Yu, X.; Zhan, Y. Study on the development law of structural fractures of Yanchang Formation in Longdong Area, Ordos Basin. Acta Sci. Nat. Univ. Pekin. 2015, 51, 1047–1058, (In Chinese with English abstract). [Google Scholar]
  116. Heng, S.; Yang, C.; Guo, Y.; Wang, C.; Wang, L. Influence of bedding planes on hydraulic fracture propagation in shale formation. J. Rock Mech. Eng. 2015, 34, 228–237, (In Chinese with English abstract). [Google Scholar]
  117. Gao, H.; Zhang, X.; He, M.; Xu, C.; Dou, L.; Zhu, G.; Li, Y. Study on evaluation of shale oil reservoir fracability based on well logging data volume. Prog. Geophys. 2018, 33, 603–612, (In Chinese with English abstract). [Google Scholar]
  118. Wang, G.; Xiong, Z.; Zhang, J. The impact of lithology differences to shale fracturing. J. Jilin Univ. 2016, 46, 1080–1089, (In Chinese with English abstract). [Google Scholar]
  119. Wang, Y.; Hou, B.; Wang, D.; Jia, Z. Features of fracture height propagation in cross-layer fracturing of shale oil reservoirs. Pet. Explor. Dev. 2021, 48, 469–479. [Google Scholar] [CrossRef]
  120. Xu, D.; Hu, R.; Gao, W.; Xia, J. Effects of laminated structure on hydraulic fracture propagation in shale. Pet. Explor. Dev. 2015, 42, 573–579. [Google Scholar] [CrossRef]
  121. Zhao, X.; Zhou, L.; Pu, X.; Jin, F.; Han, W.; Shi, Z.; Chen, C.; Jiang, W.; Guan, Q.; Xu, J.; et al. Theories, technologies and practices of lacustrine shale oil exploration and development: A case study of Paleogene Kongdian Formation in Cangdong sag, Bohai Bay Basin, China. Pet. Explor. Dev. 2022, 49, 707–718. [Google Scholar] [CrossRef]
Figure 1. The identification of multiscale laminae in shale reservoir based on the combination of multiple methods, modified after Pang et al. [25], Xu et al. [20], Wu et al. [11], and Wang et al. [18]. (A) Laminae structure and their characteristics on FMI image logs. (B) The recognition of laminae on core images and XRF scan. (C) Various laminated types represented by thin section images. (D) Various laminated types represented by SEM images.
Figure 1. The identification of multiscale laminae in shale reservoir based on the combination of multiple methods, modified after Pang et al. [25], Xu et al. [20], Wu et al. [11], and Wang et al. [18]. (A) Laminae structure and their characteristics on FMI image logs. (B) The recognition of laminae on core images and XRF scan. (C) Various laminated types represented by thin section images. (D) Various laminated types represented by SEM images.
Energies 16 01661 g001
Figure 2. Descriptive terms for lamina continuity, shape, and geometry, modified after Lazar et al. [37] and Campbell [35]. Solid yellow lines indicate continuous laminae, and dashed yellow lines indicate discontinuous laminae.
Figure 2. Descriptive terms for lamina continuity, shape, and geometry, modified after Lazar et al. [37] and Campbell [35]. Solid yellow lines indicate continuous laminae, and dashed yellow lines indicate discontinuous laminae.
Energies 16 01661 g002
Figure 3. Frequency distribution of physical properties and oil saturation of different lamina structure, modified after Ning et al. [58]. (A) and (D) Frequency distribution of porosity and oil saturation of laminated shale. (B) and (E) Frequency distribution of porosity and oil saturation of layered shale. (C) and (F) Frequency distribution of porosity and oil saturation of massive shale.
Figure 3. Frequency distribution of physical properties and oil saturation of different lamina structure, modified after Ning et al. [58]. (A) and (D) Frequency distribution of porosity and oil saturation of laminated shale. (B) and (E) Frequency distribution of porosity and oil saturation of layered shale. (C) and (F) Frequency distribution of porosity and oil saturation of massive shale.
Energies 16 01661 g003
Figure 4. The main types of reservoir space contribute to the shale porosity for different lamina structure, modified after Zhang [72].
Figure 4. The main types of reservoir space contribute to the shale porosity for different lamina structure, modified after Zhang [72].
Energies 16 01661 g004
Figure 5. Ratio of generated and expelled oil of different laminar structure shales with simulated temperature, modified after Zhang et al. [86]. The TOC, Ro, and kerogen type of laminated shale are 3.2%, 0.42%, and type I, respectively, while those of massive shale are 2.05%, 0.42%, and type II1, respectively.
Figure 5. Ratio of generated and expelled oil of different laminar structure shales with simulated temperature, modified after Zhang et al. [86]. The TOC, Ro, and kerogen type of laminated shale are 3.2%, 0.42%, and type I, respectively, while those of massive shale are 2.05%, 0.42%, and type II1, respectively.
Energies 16 01661 g005
Figure 6. Effects of pore size, pore type, and adsorption layer thickness (ta) on the proportion of adsorbed pore volume in the total space, modified after Wang et al. [107].
Figure 6. Effects of pore size, pore type, and adsorption layer thickness (ta) on the proportion of adsorbed pore volume in the total space, modified after Wang et al. [107].
Energies 16 01661 g006
Figure 7. Difference in free oil in shale with different lamina structures, modified after Sun [59].
Figure 7. Difference in free oil in shale with different lamina structures, modified after Sun [59].
Energies 16 01661 g007
Figure 8. The influence of laminar structure on shale fracability, modified after Xiong et al. [56,114]. (A,B) The effects of laminar number on shale fracability index and Poisson’s ratio. (C,D) The effects of laminar average thickness on shale fracability index and Poisson’s ratio. (E,F) The effects of laminar continuity on shale fracability index and Poisson’s ratio.
Figure 8. The influence of laminar structure on shale fracability, modified after Xiong et al. [56,114]. (A,B) The effects of laminar number on shale fracability index and Poisson’s ratio. (C,D) The effects of laminar average thickness on shale fracability index and Poisson’s ratio. (E,F) The effects of laminar continuity on shale fracability index and Poisson’s ratio.
Energies 16 01661 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xu, S.; Gou, Q. The Importance of Laminae for China Lacustrine Shale Oil Enrichment: A Review. Energies 2023, 16, 1661. https://doi.org/10.3390/en16041661

AMA Style

Xu S, Gou Q. The Importance of Laminae for China Lacustrine Shale Oil Enrichment: A Review. Energies. 2023; 16(4):1661. https://doi.org/10.3390/en16041661

Chicago/Turabian Style

Xu, Shang, and Qiyang Gou. 2023. "The Importance of Laminae for China Lacustrine Shale Oil Enrichment: A Review" Energies 16, no. 4: 1661. https://doi.org/10.3390/en16041661

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop