Abstract

Exploring the tightness mechanism through a quantitative analysis of the pore evolution process is the research hotspot of tight oil reservoirs. The physical characteristics of Chang 6 (Ch-6) sandstones in the western Jiyuan area have the typical features of a tight oil reservoir. Based on the reservoir physical property, lithological characteristics, diagenetic types and sequence, and burial and thermal evolution history, this study analyzes the factors leading to reservoir tightness and establishes the model of the pore evolution process. The results show that the sedimentary microfacies type controls the reservoir detrital material and further affects its physical properties. The high content of feldspar and rock fragments and the fine grain size are the material cause for the reservoir tightness. The sandstones of the main underwater distributary channel are the dominant sedimentary bodies for the development of a high-quality reservoir. In terms of diagenesis, compaction is the primary cause for reservoir tightness, and the porosity reduction by cementation is weaker than that by compaction. Meanwhile, the quantitative calculation results indicate that the porosity losses by compaction, carbonate cementation, kaolinite cementation, chlorite coatings, and siliceous cementation are 23.5, 3.1, 3.8, 3.0, and 0.8%, respectively. In addition, dissolution is significant to improve the reservoir physical property, and the increase of dissolved porosity is around 3.2%. More significantly, this study uses a detailed and systematic method for analyzing the tightness mechanism and the pore evolution process of the Ch-6 sandstones in the western Jiyuan area, Ordos Basin, China.
1. Introduction
Unconventional oil and gas are important components of hydrocarbon resources, including shale gas, shale oil, tight sandstone oil, natural gas hydrates, etc.1−5 Tight sandstone oil comes from sandstone reservoirs with porosity less than 10%, air permeability less than 1 mD, or in situ matrix permeability less than 0.1 mD.6 The sandstones of the Upper Triassic Yanchang formation in the Ordos basin are characterized by low porosity, ultralow permeability, and abundant oil resources.7−10 In the Yanchang formation, Chang 6 (Ch-6) has huge potential for exploration and development of tight oil, with reserves of more than 100 million tons in the Jiyuan, Xifeng, and Huaqing areas.11−15
At present, quantitative analysis of the reservoir pore evolution is the research hotspot in the study of reservoir and diagenesis. In 1987, David firstly established the calculation method of porosity evolution using thin-section analysis data and achieved good results in the calculation of the compaction porosity loss of Jurassic feldspathic quartz sandstones in Utah and Ordovician quartz sandstones in Oklahoma.16 Later, Ehrenberg demonstrated the quantitative simulation method of porosity evolution and proposed the useful supplements of the compaction porosity loss calculation method, statistical method of thin sections, effects of particle solution on compaction porosity loss, and repeatability of the calculated results.17 In addition, the studies on the Ventura Basin in California by Wilson and that on the Bermejo Basin in Argentina by Damanti have achieved effective applications.18,19 In terms of Chinese research studies, many scholars have done many works in the quantitative analysis of the reservoir pore evolution and obtained productive results in the Ordos Basin, Bohai Bay Basin, Junggar Basin, and Songliao Basin.20−25
Previous studies focused more on the overall effect of compaction and cementation but ignored the individual effects of various types of cementation and the “specific process” of reservoir tightness.25−27 At the same time, from the simulation method of pore evolution based on thin-section data, it is difficult to accurately evaluate the dissolved contents of each component and the variation of rock skeleton volume in the diagenetic process. Such statistical errors need to be inhibited through the method of sample screening and data processing.24
Above all, this paper uses a detailed and systematic method for studying the reservoir tightness mechanism via the analysis of the porosity evolution process of the Ch-6 reservoir in the western Jiyuan (WJY) area, hoping that it could provide a scientific basis for the relative studies of tight oil reservoirs in other areas.
2. Geological Setting
The Ordos Basin is a polycyclic craton margin basin in North China, which is characterized by stable subsidence and multiple migrations of depocenters, and it is composed of the Precambrian basement and Phanerozoic sediments.28,29 The oil and gas resources of the Ordos Basin are abundant, and the petroleum geology conditions are relatively simple.30,31 The internal structure of the basin is relatively simple, and the strata are relatively gentle with a dip angle of less than 1°, but faults and thrust structures abundantly exist in the western margin due to the later influences of the Yanshan movement.14 According to the nature of the basin basement, the structure form, and characteristics, it can be divided into six first-level tectonic units: the Yimeng uplift, the Weibei uplift, the west margin thrust belt, the Jinxi flexure belt, the Tianhuan depression, and the Shanbei slop.32 The western Jiyuan (WJY) area, which is located in the western area of the Ordos Basin, spans the Tianhuan depression (Figure 1a,b) and presents a gentle monoclinal structure sloping westward. In the sedimentary period of Yanchang formation, the basin is a half-graben-like sag, extending along the south and north directions, with a steep and narrow western flank and a moderate eastern flank.8 A complete continental river–delta–lake depositional system is developed in the Triassic Yanchang formation of the Ordos Basin, which continuously and stably presents the three stages of lake-transgression to lake-flooding to lake regression. Accordingly, the Yanchang formation records the whole process of the formation, development, prosperity, decline, and extinction of the lake basin with a 12 Ma time span, which can be subdivided into 10 areas represented as Ch-1 (at the top) to Ch-10 (at the bottom).32 Ch-6 is the regressive deposit after the heyday of the Ch-7 lake basin, which develops the deep lake, semideep lake, and delta facies from the basin center to the margin. The thickness of the Ch-6 formation is 80–110 m (Figure 2), which is comfortably in contact with the Ch-7 formation; the typical delta deposits are mainly composed of gray sandstone, dark mudstone, and thin coal seam.33 In the WJY area, the underwater distributary channels are well developed in the delta front and the distributary channels are partly developed in the delta plain, which are controlled by the provenance of northwestern Alxa ancient land.34
Figure 1.
(a) Ordos Basin tectonic units and location of the study area. (b) Map showing the distribution of CH-61 sand bodies in the WJY area.
Figure 2.
Lithology section and depositional environment of the Upper Triassic Yanchang formation in the Ordos Basin, China.
3. Materials and Methods
All kinds of data used in this study were collected from the core samples of the Ch-6 sandstone reservoirs in the WJY area. First, the data on porosity, permeability, grain size, and the mercury pressure test were collected from the Changqing Oilfield Company. Based on the data on the physical properties of 414 samples and the mercury pressure data of 12 samples, this study analyzed the pore structure and pore-throat features of the Ch-6 tight oil reservoirs. At the same time, data on the grain size obtained from nine samples were utilized to assess particle sizes and calculate the sorting coefficient (S0) as well as the initial porosity of the Ch-6 reservoirs according to the method published by Scherer.35
In addition, 62 samples were collected from 43 Wells for microscopic petrology and pore-throat identification in the Key Laboratory of Petroleum Resources Research, Institute of Geology and Geophysics, Chinese Academy of Sciences. By observing 152 thin sections, the statistical analysis was performed, which includes mineral type and content, the degree of sorting and roundness, pore type, current thin porosity, the content of dissolved pores, the content of original intergranular pores, and cements content. The thin sections were impregnated with red and blue epoxy resin to highlight the pore spaces, and some of them were stained with Alizarin Red S and K-ferricyanide for carbonate mineral identification. The mineral composition and pore space were observed with Axioskop 40 polarizing microscope manufactured by Zeiss. An optical collection system was used to quantify thin-section porosity and diagenetic cements content by counting at least 350 points of each thin section. The method of point counting and proportion counting was mainly utilized in the quantitative calculation of detrital components and all kinds of pore content.16,17 Moreover, 11 freshly broken chips of reservoir samples were prepared, and the MERLIN field emission scanning electron microscope (FE-SEM) equipped with the energy-dispersive X-ray (EDS) system manufactured by Zeiss was used to carry out a more detailed observation on the morphology of authigenic clay minerals and intergranular cement. On the basis of thin-section observation, the diagenetic stages of Ch-6 in the WJY area were identified according to the China Petroleum Standard of the Division of Diagenetic Stage in Clastic Rocks (SY/T 5477-2003). At last, mathematical statistics were used in the quantitative calculation of porosity in different diagenetic stages following the methods published by Lai et al.36
The carbon and oxygen isotopes of 45 sandstone samples were measured in the Key Laboratory of Petroleum Resources Research, Institute of Geology and Geophysics, Chinese Academy of Sciences. The specific process is to carry out the accurate measurement of carbon and oxygen stable isotopes of CO2 that were released from the dissolution of carbonate materials by H3PO4 with a MAT-252 isotope mass spectrometer. The measurement accuracy is less than 0.5‰, and all isotope data were calibrated with standard samples including NBS-18, GBW04405, and GBW04406. Ultimately, the data were presented relative to the Vienna PeeDee Belemnite (V-PDB) and in the δ notation.
4. Results
4.1. Reservoir Lithology
Through observation and statistical analysis of more than 150 thin sections, the Ch-6 reservoirs in the WJY area were found to be mainly composed of lithic arkose, feldspathic litharenite, and a small amount of arkose according to Folk’s classification scheme37 (Figure 3a). The detrital compositions of reservoir sandstones are dominated by quartz, ranging from 11.29 to 46.67% with an average of 29.83%, feldspar, ranging from 7.17 to 55.17% with an average of 35.17%, and rock fragments, ranging from 1.3 to 34.2% with an average of 10.81%. On comparing the fraction of rock fragments from a different origin, the content of metamorphic rock fragments was found to be the highest (from 1.13 to 15.7% with an average of 5.97%), followed by sedimentary and volcanic rock fragments (from 0 to 9.47% with an average of 2.47%, from 0.17 to 9.03% with an average of 2.37%, respectively) (Figure 3b). In terms of grain size, the Ch-6 reservoir sandstones are mainly composed of fine sandstone and siltstone (Figure 4a). In addition, the shape of grains is mainly subrounded to subangular with moderately sorted composition (Figure 4b).
Figure 3.
Rock compositions of the Ch-6 sandstones in the WJY area, Ordos Basin. (a) Ternary diagram showing framework grain compositions. (b) Histogram displaying the compositions of rock fragments. Q, Quartz; F, feldspar; R, rock fragments; I, quartz arenite; II, subarkose; III, sublitharenite; IV, arkose; V, lithic arkose; VI, feldspathic litharenite; and VII, litharenite.
Figure 4.
Histograms of (a) grain size and (b) sorting characteristics of Ch-6 in the WJY area, Ordos Basin.
4.2. Reservoir Properties
4.2.1. Porosity and Permeability
The porosity of the Ch-6 sandstones in the WJY area mainly ranges from 6 to 12% with an average of 8.59% and the permeability mainly ranges from 0.1 to 0.5 mD with an average of 0.32 mD (Figure 5). Therefore, these sandstones have the typical characteristics of low porosity and permeability, which belong to the category of tight oil reservoirs defined by Zou et al., Yao et al., and Kuang et al.3,7,38 Moreover, there is a roughly exponential correlation (R2 = 0.4686) between porosity and permeability, which shows that the permeability is mainly controlled by the development of the connected pores (Figure 6).
Figure 5.
Distribution of (a) porosity and (b) permeability of the Ch-6 reservoirs in the WJY area, Ordos Basin.
Figure 6.

Relationship between the porosity and permeability of the Ch-6 reservoirs in the WJY area, Ordos Basin.
4.2.2. Pore Types and Pore-Throat Characteristics
Porosity determines the reservoir capacity, and permeability is an important factor for reservoir productivity.39,40 The pore types of the Ch-6 reservoirs in the WJY area are mainly original intergranular pores, feldspar dissolved pores, rock fragments dissolved pores, intercrystalline pores, and microcracks, the thin-section porosity of which are 1.60, 0.84, 0.16, 0.09, and 0.03% (Figure 7a), respectively. Therefore, the original intergranular pore and dissolved pores account for 95.58% of the total thin-section porosity (Figure 7b).
Figure 7.
Pore type distribution diagram of the Ch-6 reservoir in the WJY area, Ordos Basin. (a) Diagram showing the percentages of various pore types. (b) Diagram showing the total thin-section porosity and the different types of porosities.
Mercury injection analysis is an important method to study the reservoir pore structure, which can quantitatively reflect the microstructure characteristics of the reservoir pore throat. Based on the analysis and calculation of typical capillary pressure curves and pore-throat radius value distribution of the Ch-6 reservoir (Figure 8), it could be found that the mean displacement pressure ranges from 0.21 to 2.53 MPa with an average of 1.22 MPa, the median throat radius value ranges from 0.06 to 0.51 μm with an average of 0.24 μm, the measured porosity ranges from 6.4 to 16.3% with an average of 9.94%, and the measured permeability ranges from 0.06 to 0.68 mD with an average of 0.24 mD. It is found by further analysis that the measured porosity and permeability are positively correlated with the median throat radius values and negatively correlated with the mean displacement pressure. However, it must be pointed out that due to the limitations of experimental methods, mercury injection analysis could only characterize the characteristics of the partial pore system but could not characterize the full-scale pore-throat structure of sandstones.41,42 Hence, the experimental data here mainly describe the pore structure features above the mesopores in sandstones.
Figure 8.
Typical capillary pressure curves and pore-throat radius distribution of the Ch-6 reservoirs in the WJY area, Ordos Basin. (A) Well C193, 2242.72 m; (B) Well Y60, 2549.81 m; (C) Well Y27, 2348.40 m; and (D) Well Y94, 2739.48 m.
4.3. Diagenetic Types
4.3.1. Compaction
Compaction is one of the significant factors leading to the tightness of the Ch-6 reservoirs in the WJY area and could be divided into two types: mechanical compaction and chemical compaction. By observing the thin sections, it could be seen that the contact relation between the particles is mainly a linear contact, which preserves the original intergranular pores (Figure 9a) at the early stage of burial. As the burial depth increases and the paleogeotemperature exceeds 70 °C,43 chemical compaction starts and the contact relation between the particles tends to a linear-suture contact (Figure 9b); also some mica particles undergo a plastic deformation (Figure 9c), where the original intergranular pores are relatively less.
Figure 9.
Microscopic characteristics of the Ch-6 sandstones in the WJY area, Ordos Basin. (a) Original intergranular pores (yellow arrow); the brown chlorite coatings on the surface of the grains after being soaked by oil (green arrow); Well F11, 2368.33 m, plane-polarized light (PPL). (b) Grains exhibit a linear contact due to compaction (blue arrow), and some of them have microcracks (yellow arrow); Well Y83, 2492.42 m, PPL. (c) Plastic deformation of mica fragments due to compaction, which are in close contact with the rigid grains (yellow arrow); Well H35, 2515.63 m, PPL. (d) Intergranular pores (green arrow) and feldspar dissolved pores (blue arrow) are cemented by ferrocalcite; Well H307, 2736.84 m, PPL; (e) Intergranular pores are cemented by ferrocalcite (yellow arrow); Well H315, 2580.75 m, PPL. (f) Secondary growth on the surface of terrigenous quartz grains (yellow arrow); Well H306, 2739.40 m, PPL. (g) Authigenic quartz with a good crystal morphology in dissolution pores (yellow arrow); Well Y76, 2379.90 m, SEM. (h) Microcrystalline chlorites, which grow vertically on the surface of the grains in the intergranular pore, form coatings on the surface of the grains (yellow arrow); Well F11, 2369.17 m, PPL. (i) Chlorite with the crystalline morphology of a petaline shape on the surface of the grains (yellow arrow); Well Y28, 2423.00 m, SEM. (j) Kaolinite in feldspar dissolved pores (yellow arrow); Well C335, 2303.36 m, PPL. (k) Authigenic kaolinite grains with accordion morphologies occur in pore spaces (yellow arrow); Well Y28, 2442.78 m, SEM. (l) Illite occurs as pore bridges with the crystalline morphology of a needlelike and hairy structure (yellow arrow); Well Y76, 2385.80 m, SEM. (m) Feldspar dissolved pores (yellow arrow) and moldic pores (red arrow); Well Y76, 2338.63 m, PPL. (n) Intergranular pores (green arrow) and debris dissolved pores (yellow arrow); Well C335, 2303.36 m, PPL. (o) Feldspar grains are dissolved (yellow arrow) and authigenic kaolinite is developed in the dissolved pores (red arrow); Well Y28, 2448.23 m, SEM.
4.3.2. Cementation
Cementation is an important consolidation diagenesis due to the precipitation of authigenic minerals in the pores, which lasts throughout the diagenesis process except for compaction. By observing the thin sections, it was seen that the main types of cementation of the Ch-6 reservoirs are composed of carbonate cementation, siliceous cementation, and clay minerals cementation.
4.3.2.1. Carbonate Cementation
Carbonate cement is one of the most common authigenic minerals in clastic rock reservoirs that are continuously superimposed in the vertical direction with multiple argillaceous interlayers. By observing the thin sections of the Ch-6 reservoir sandstones in the WJY area, it was seen that the carbonate cements are dominated by ferrocalcite (Figure 9d,e) with an average content of 4.10%. They are mainly distributed in the intergranular pores and/or the intragranular dissolution pores of feldspar and show up as deep red after being stained. These features indicate that the later dissolution of feldspars is likely to provide the necessary calcium ions for their precipitation.
4.3.2.2. Siliceous Cementation
Compared with carbonate cement, siliceous cement is limited and its average content is 1.1%, which is characterized by terrigenous quartz overgrowth and authigenic quartz crystals in dissolved pores (Figure 9f,g).
4.3.2.3. Clay Cementation
The authigenic clay minerals of the Ch-6 reservoir sandstones in the WJY area are primarily composed of chlorite, kaolinite, and illite. Chlorites primarily appear as a grain coating, and this form of pore-filling cement is rare; the average content of chlorite is around 4.9%. Authigenic kaolinite fills the pores in the form of interstitial material (Figure 9j) and exhibits accordion morphologies (Figure 9k), and the average content of kaolinite is around 3.9%. Illite occurs as pore bridges with the crystalline morphology of a needlelike and hairy structure according to the SEM observation (Figure 9l), and the average content of illite is less than 1%.
4.3.3. Dissolution
The main dissolved grains are detrital feldspars and rock fragments, usually along the grain edges and cleavage fractures to form intragranular and intergranular dissolved pores and little moldic pores (Figure 9m–o). However, the dissolution of pore-filling carbonate cements is relatively rare. Meanwhile, the reservoir property would be improved effectively when the dissolved matters are moved out of the pore or throat. However, the permeability of the reservoir would be greatly reduced when these matters precipitate in the intergranular pore in the form of authigenic clay minerals. The results in this study suggest that dissolution is the most significant factor to improve the physical properties of the Ch-6 reservoirs in the WJY area, and the average content of the dissolved pores is about 4.2%.
5. Discussion
5.1. Diagenetic Stage and Sequence
On the basis of the aforementioned sedimentary characteristics and predecessor research results,44 the sedimentary environment of Ch-6 formation in the WJY area is identified as a fresh-brackish inland lake basin. The dividing standard of the diagenetic stage and sequence is according to the China Petroleum Standard of the Division of Diagenetic Stage in Clastic Rocks (SY/T 5477-2003).
The maximum buried depth of the Ch-6 reservoirs in the WJY area ranges from 2500 to 2700 m, corresponding to the maximum paleotemperature ranging from 100 to 125 °C with an ancient land surface temperature of 15 °C and a paleogeothermal gradient of about 3.5–4.0 °C/100 m.44 The vitrinite reflectance (Ro) of the Ch-6 mudstone mainly ranges from 0.5 to 0.9% and a small number of samples exceed to 1.2%. The smectite content of mixed layer I/S is under 15%; the homogenization temperature of fluid inclusions in the margin of quartz secondary enlargement and the microcracks ranges from 70 to 170 °C, which holds two distinct temperature peak areas of 80 to 100 and 120 to 130 °C.10 These data indicate that the Ch-6 tight oil reservoir is at the mesogenetic A stage. Furthermore, the petrographic characteristics of the Ch-6 tight oil reservoirs (e.g., point to linear contact between grains, the types of dissolved minerals, which are mainly composed of feldspar and detritus, etc.) also indicate that they belong to the mesogenetic A stage (Figure 10).
Figure 10.
Classification symbols of the CH-6 diagenetic evolution stage in the WJY area, Ordos Basin.
Additionally, the δ18O value of carbonate cements is usually used to calculate the paleotemperature and as a geological thermometer for determining the ambient temperatures, which can evaluate the environmental conditions and the diagenetic stages of cementation. In this study, the mature theoretical equation for geological thermometers proposed by Shackleton et al. is used to calculate the paleotemperature (T).45 The paleotemperature is actually the forming temperature of carbonate cements and it is calculated using the following equation.
| 1 |
where the δ18O uses the PDB standard; δ18Ow is the oxygen isotope values of geological fluids medium at the time carbonate cements formation, and the value of δ18Ow is selected as −2.5‰ according to the overall geological environment at the time of the Ch-6 diagenesis period.
The calculation results show that the paleotemperature of carbonate cements in the Ch-6 reservoirs ranges from 95 to 110 °C (Table 1), which corresponds to the mesogenetic A stage.
Table 1. Carbon and Oxygen Isotopic Data of Ch-6 Carbonate Cements in the WJY Area.
| no. | well name | depth (m) | δ13CPDB (‰) | δ18OPDB (‰) | value of Z | paleotemperature (°C) |
|---|---|---|---|---|---|---|
| 1 | C118 | 2266.5 | –1.913 | –21.547 | 112.65 | 101.78 |
| 2 | C18 | 2153.9 | –0.788 | –20.972 | 115.24 | 97.62 |
| 3 | C18 | 2145.5 | 1.09 | –15.022 | 122.05 | 58.47 |
| 4 | C18 | 2184.86 | –0.449 | –20.594 | 116.12 | 94.92 |
| 5 | F10 | 2216.9 | –0.763 | –22.221 | 114.67 | 106.74 |
| 6 | H132 | 2336.15 | –1.909 | –19.55 | 113.65 | 87.62 |
| 7 | H132 | 2368.25 | 1.361 | –20.877 | 119.69 | 96.94 |
| 8 | H269 | 2385.8 | –1.05 | –18.512 | 115.93 | 80.58 |
| 9 | H269 | 2396.3 | 0.325 | –21.48 | 117.27 | 101.29 |
| 10 | H269 | 2366 | –0.989 | –22.307 | 114.17 | 107.38 |
| 11 | H269 | 2408.7 | –0.289 | –23.109 | 115.2 | 113.41 |
| 12 | H269 | 2419.27 | 0.204 | –20.341 | 117.59 | 93.13 |
| 13 | H48 | 2480.1 | –0.84 | –21.819 | 114.71 | 103.77 |
| 14 | H83 | 2455.7 | –1.674 | –20.687 | 113.57 | 95.58 |
| 15 | H83 | 2662 | –8.742 | –21.48 | 98.7 | 101.29 |
| 16 | Y114 | 2273.6 | –2.202 | –21.161 | 112.25 | 98.98 |
| 17 | Y115 | 2298.3 | 0.184 | –14.506 | 120.45 | 55.41 |
| 18 | Y115 | 2313.8 | 0.399 | –14.511 | 120.89 | 55.44 |
| 19 | Y115 | 2382.05 | –2.428 | –18.963 | 112.88 | 83.61 |
| 20 | Y117 | 2349.4 | –1.751 | –19.647 | 113.93 | 88.29 |
| 21 | Y117 | 2351.2 | –1.348 | –14.11 | 117.51 | 53.1 |
| 22 | Y117 | 2353.7 | –1.422 | –17.021 | 115.91 | 70.83 |
| 23 | Y123 | 2376.1 | –3.806 | –21.038 | 109.03 | 98.09 |
| 24 | Y28 | 2440.8 | –1.995 | –21.482 | 112.52 | 101.3 |
| 25 | Y28 | 2443.1 | –1.262 | –22.351 | 113.58 | 107.7 |
| 26 | Y28 | 2448.28 | –2.138 | –21.219 | 112.35 | 99.4 |
| 27 | Y41 | 2506.8 | –1.787 | –17.555 | 114.9 | 74.27 |
| 28 | Y41 | 2508.3 | –0.473 | –15.095 | 118.81 | 58.91 |
| 29 | Y41 | 2547.7 | –4.491 | –21.982 | 107.16 | 104.97 |
| 30 | Y41 | 2556.65 | –3.48 | –21.208 | 109.61 | 99.32 |
| 31 | Y41 | 2571.4 | –2.566 | –22.008 | 111.08 | 105.16 |
| 32 | Y71 | 2370.1 | –1.67 | –19.746 | 114.05 | 88.97 |
| 33 | Y71 | 2415.9 | –0.157 | –20.799 | 116.62 | 96.38 |
| 34 | Y73 | 2375.3 | 0.056 | –20.652 | 117.13 | 95.33 |
| 35 | Y76 | 2345.7 | –0.27 | –25.53 | 114.03 | 132.4 |
| 36 | Y76 | 2337.7 | –0.312 | –23.855 | 114.78 | 119.13 |
| 37 | Y76 | 2339.7 | –0.141 | –24.115 | 115 | 121.16 |
| 38 | Y76 | 2385.62 | 2.317 | –24.793 | 119.7 | 126.49 |
| 39 | Y76 | 2385.62 | –1.218 | –20.397 | 114.65 | 93.53 |
| 40 | Y76 | 2378.95 | –0.224 | –21.617 | 116.08 | 102.29 |
| 41 | Y76 | 2372.6 | 2.127 | –24.298 | 119.56 | 122.59 |
| 42 | Y76 | 2378.95 | 2.031 | –24.221 | 119.4 | 121.98 |
| 43 | Y76 | 2379.9 | 0.147 | –20.962 | 117.16 | 97.55 |
| 44 | Y76 | 2385.8 | –0.769 | –21.365 | 115.09 | 100.45 |
| 45 | Y76 | 2393.6 | –3.002 | –19.711 | 111.34 | 88.73 |
Comprehending the observation of diagenesis and textural relationships of authigenic minerals with the results of the diagenetic stage analysis and isotopic data, the diagenetic evolution sequence of the Ch-6 reservoirs in the WJY area is as follows: mechanical compaction, formation of chlorite coating, formation of quartz overgrowth, feldspar dissolution, formation of authigenic kaolinite, formation of authigenic quartz, conversion of kaolinite to illite, precipitation of ferrocalcite, and, finally, hydrocarbon emplacement.
5.2. Controlling Factors of Reservoir Quality
5.2.1. Sedimentary Conditions Determine the Material Basis of the Reservoir
Sedimentary conditions control the reservoir material basis. Fundamentally, sedimentary facies control the microscopic distribution of reservoirs and affect the basic physical properties of the reservoir. In this study, the physical parameters of sand bodies of different sedimentary microfacies were calculated. The results show that the average porosity and permeability of the main channel sandstones in the delta plain are 6.61% and 0.08 mD, respectively; the average porosity and permeability of the channel margin sandstones in the delta plain are 5.04% and 0.05 mD, respectively; the average porosity and permeability of the main channel sandstones in the delta front are 9.53% and 0.32 mD, respectively; the average porosity and permeability of the channel margin sandstones in the delta front are 4.93% and 0.13 mD, respectively; the average porosity and permeability of interdistributary bay sands is 3.48%, 0.05 mD, respectively (Table 2). In consequence, the main channel sandstone of the underwater distributary channel in the delta front is the dominant sedimentary body for the relatively high-quality reservoir development.
Table 2. Physical Parameters of Sand Bodies with Different Sedimentary Microfacies.
| distributary
channel |
underwater distributary
channel |
||||
|---|---|---|---|---|---|
| physical parameter | main channel sandstones | channel margin sandstones | main channel sandstones | channel margin sandstones | interdistributary bay sandstones |
| porosity (%) | 6.61 | 5.04 | 9.53 | 4.93 | 3.48 |
| permeability (mD) | 0.08 | 0.05 | 0.32 | 0.13 | 0.05 |
Moreover, petrological characteristics (e.g., lithology, composition, grain size, sorting, and roundness) also affect reservoir physical properties, while the detailed impact mechanism is comprehensive and multifactorial.46,47 Therefore, it is difficult to find the correlation between a single factor and reservoir physical characteristics. This study attempts to use the mathematical statistics methods and pick several influence factors that are well correlated with reservoir physical properties (Figure 11 and 12). The result shows that the average particle size is positively associated with thin-section porosity (Figure 12). Quartz content is positively associated with intergranular pore content. The total content of feldspar and rock fragment is positively associated with the thin-section porosity of dissolved pores (Figure 11). Further, according to the lithologic characteristics of the reservoir, the high content of feldspar and rock fragment as well as the fine particle size are the material causes for the weak compaction capability of the reservoirs and partly the reasons for the reservoir tightness.
Figure 11.
Correlation between intergranular pore content and quartz content; the correlation between thin-section porosity of dissolved pores and the total content of feldspar and rock fragments.
Figure 12.

Correlation between thin-section porosity and the average particle size.
5.2.2. Reservoir Transformation due to Diagenesis
As mentioned above, sedimentary conditions control the petrologic basis and spatial distribution of reservoirs. Furthermore, after deposition, as the burial depth increases, the diagenetic conditions (temperature, pressure, pH, and fluid solubility) change, and the diagenetic features show obvious diversity in each diagenetic stage. Accordingly, the diagenesis diversity can affect the pore evolution and physical characteristics of the reservoirs.
To assess the relative importance of compaction and cementation in the ultimate reservoir quality, the diagram of Houseknecht et al. is applied in the study16 (Figure 13). The diagram shows that more data points cluster in the lower-left portion of the diagram, which indicates that the reservoirs are mainly controlled by compaction in the process of tightness and the effect of cementation is relatively weak.
Figure 13.

Intergranular volume vs cements volume for the Ch-6 sandstones in the WJY area, Ordos Basin.
In addition, the effects of different diageneses on reservoir properties are studied through correlation analysis. The result suggests that the content of carbonate cements, kaolinite fillings, chlorite coatings, and siliceous cements are all negatively associated with the measured porosity (Figure 14), which indicates that multiple cementations are an important cause of reservoir densification.
Figure 14.
Correlation between different cements with measured porosity.
5.3. Quantitative Analysis of the Pore Evolution Process
According to research requirements, original porosity (OP) is obtained as follows. The original porosity (OP) could be calculated with eq 2, where particle diameters of P25 and P75 represent the corresponding cumulative volume fraction of 25 and 75%, respectively. The Trask sorting coefficient (S0) can be calculated with eq 3. The calculated results show that S0 is 1.32 and the original porosity (OP) of the Ch-6 reservoirs in the WJY area is 38.2%.
| 2 |
| 3 |
In addition, it is significantly necessary to analyze the quantitative influence of various diagenesis on reservoirs. Based on the burial history, paleogeothermal history44 (Figure 15), porosity evolution progress, and the aforementioned diagenetic evolution sequence of Ch-6 in the WJY area, analyzing the reservoir changes due to the various diagenesis types, this study established the quantitative progress of pores evolution and divided the progress into seven main stages (Figure 16).
Figure 15.
Burial and paleogeothermal history of the Ch-6 reservoir in the WJY area, Ordos Basin.
Figure 16.
Model diagram of the Ch-6 reservoir tightness process and pore evolution in the WJY area, Ordos Basin.
First, in the early stages of burial and diagenesis, various diagenesis had not yet begun to transform the reservoirs. The reservoir space was mainly composed of primary intergranular pores and the contact type between particles was point-contact. As calculated above, the optional porosity was 38.2%.
Second, in the compaction porosity loss stage, the burial depth and paleogeotemperature increased to around 1500 m and 65 °C, respectively. Under the gravitational pressure of overlying sediments and tectonic stress, the grains rearranged and intergranular water was discharged, and the contact type between the grains changed to a point-linear contact and the porosity was gradually lost. In fact, compaction occurs throughout the diagenesis process and it is not a specific diagenetic action at a certain diagenetic stage. Therefore, to reflect the changing degree of the reservoir by compaction, this study considers it as an influencing parameter for quantitative analysis. The observation and statistic results of thin sections show that the intergranular volume (IGV) is about 19.2%. Based on the calculation of eq 4, the compaction porosity loss (COPL) is 23.5%. In addition, when the burial depth was around 1500 m and the paleogeotemperature was 65 °C, a small amount of early carbonate cements and clay membrane of chlorite were developed in the reservoirs. However, these two types of diagenesis have little effect on the reservoir in the early diagenetic stage, and the study does not conduct the quantitative analysis. Therefore, under the transformation of compaction, the reservoir pore type was residual primary intergranular pore and the residual porosity was about 14.7%.
| 4 |
Third, in the dissolved porosity increase stage, the burial depth and paleogeotemperature increased to around 1700 m and 85 °C, respectively. The soluble components were dissolved due to the change of the diagenetic environment. Similar to compaction, dissolution also occurred throughout the diagenesis process and the study considers it as an influencing parameter for quantitative analysis. The results of the analysis indicate that the dissolution minerals are mainly unstable minerals such as early carbonate cementation, feldspar, and rock fragments. The pore types are residual primary intergranular pores and intergranular dissolved pores during this stage. In addition, with the increase of compaction, the contact relation between grains changes from point to point-line. The observation and statistic results of thin sections show that the dissolved porosity (DP) is about 4.2%. Based on the calculation as in eq 5, the dissolved porosity increase (DPI) was around 3.2% and the residual porosity of the reservoir was 17.9%.
| 5 |
Fourth, in the chlorite cementation porosity loss stage, the burial depth and paleogeotemperature increased to around 1900 m and 100 °C, respectively. The authigenic chlorite of the Ch-6 reservoirs is mainly developed in the form of a grain coating. The observation and statistic results of the thin sections show that the content of chlorite cements (CHCEM) is about 4.9%. Based on the calculation as in eq 6, chlorite cementation porosity loss (CHCEPL) was around 3.8% and the residual porosity of the reservoir was 14.1%.
| 6 |
Fifth, in the kaolinite cementation porosity loss stage, the burial depth and paleogeotemperature increased to around 2250 m and 125 °C, respectively. Authigenic kaolinite is one of the common authigenic clay minerals in the Ch-6 reservoir of the WJY area, which is mainly filled in the dissolved pores of feldspar and residual original intergranular pores and closely related to the unstable aluminosilicate minerals (such as feldspar). The observation and statistic results of the thin sections show that the content of kaolinite cements (KCEM) is about 3.9%. Based on the calculation as in eq 7, the kaolinite cementation porosity loss (KCEPL) was 3.0% and the residual porosity of the reservoir was 11.1%.
| 7 |
Finally, in the carbonate cementation porosity loss and siliceous cementation porosity loss stage, the burial depth of the Ch-6 reservoir increased to the maximum burial depth of about 2700 m, and then the stratum was continually uplifted to the present burial depth of about 2100 m. The medium-term carbonate cements are mainly sparry calcite or ferrocalcite, but the late carbonate cements are mainly ankerite and to a lesser extent ferrocalcite in the Ch-6 reservoirs of the WJY area. The early carbonate cements are scarcely observed due to the dissolution. The observation and statistic results of the thin sections show that the content of carbonate cements (CACEM) is about 4.1%. Based on the calculation as in eq 8, the carbonate cementation porosity loss (CACEPL) was 3.1% and the residual porosity was around 8.0% after the carbonate cementation. In addition, the main types of siliceous cements in the Ch-6 reservoirs of the WJY area are terrigenous quartz overgrowth and authigenic quartz crystals in dissolved pores. The observation and statistic results of the thin sections show that the content of siliceous cements (SCEM) is about 1.1%. Based on the calculation as in eq 9, the siliceous cementation porosity loss (SCEPL) was 0.8% and the residual porosity of the reservoir was about 7.2%. Inescapably, the illite was rarely developed in the reservoirs in this stage and had little effect on the reservoir. Therefore, the study ignores it as an influencing parameter and does not conduct a quantitative analysis of illite cementation.
| 8 |
| 9 |
Above all, on account of the quantitative analysis of porosity evolution, compaction is the primary cause leading to reservoir tightness. The other diagenetic factors leading to reservoir tightness are cementation of chlorite, carbonate, kaolinite, silica, and illite, ranked from a strong to a weak degree of influence on reservoir tightness. In addition, dissolution is the key factor to improve the physical property of the reservoir. The tightness process and pore evolution model of the Ch-6 reservoir in the WJY area are shown in Figure 16 (Table 3).
Table 3. Meaning of the Abbreviations in the Above Equations.
| OP | original porosity: the porosity before all kinds of diagenesis |
| COPL | compaction porosity loss: the percentage of the pore volume reduced after compaction in the apparent original rock volume |
| IGV | intergranular volume: the porosity that has undergone compaction but without dissolution |
| DPI | dissolved porosity increase: the percentage of pore volume resulting from dissolution in the apparent volume before all kinds of diagenesis |
| DP | dissolved porosity: the percentage of pore volume resulting from dissolution in the apparent volume at present |
| CHEPL | chlorite cementation porosity loss: the percentage of reduced pore volume due to chlorite cementation in the apparent volume before all kinds of diagenesis |
| CHCEM | chlorite cements content: the percentage of chlorite cements content in the rock volume |
| KCEPL | kaolinite cementation porosity loss: the percentage of reduced pore volume due to kaolinite cementation in the apparent volume before all kinds of diagenesis |
| KCEM | kaolinite cements content: the percentage of kaolinite cements content in the rock volume |
| CACEPL | carbonate cementation porosity loss: the percentage of reduced pore volume due to carbonate cementation in the apparent volume before all kinds of diagenesis |
| CACEM | carbonate cements content: the percentage of carbonate cements content in the rock volume |
| SCEPL | silica cementation porosity loss: the percentage of reduced pore volume due to silica cementation in the apparent volume before all kinds of diagenesis |
| SCEM | silica cements content: the percentage of silica cements content in the rock volume |
6. Conclusions
Sedimentary facies types primarily controlled the material basis of the Ch-6 reservoirs in the WJY area and further affected their physical properties. The high content of feldspar and rock fragments and the fine grain size are the material causes for the reservoir tightness. The main channel sandstone of the underwater distributary channel is the dominant sedimentary facies type for the high-quality reservoir development.
The diagenetic stage of the reservoir Ch-6 in the WJY area is identified as the late stage of eodiagenesis A to the early stage of mesodiagenesis B. The main diagenetic factors that lead to the reservoir tightness include compaction and the cementation of carbonate, chlorite kaolinite, and silica, the porosity loss of which are 23.5, 3.1, 3.8, 3.0, and 0.8%, respectively. The dissolved porosity increase is around 3.2%. Therefore, compaction and the cementation of chlorite, kaolinite, and carbonate are the key factors to damage the physical property of the reservoir. The dissolution is the most significant factor to improve reservoir physical property.
Acknowledgments
This study was jointly supported by funds from the National Natural Science Foundation of China (Grant Nos. 41772142; 41802160) and the National Construction of High-Quality University Projects of Graduates from China Scholarship Council (Grant No. 201904910319).
The authors declare no competing financial interest.
References
- Zou C.-n.; Tao S.-z.; Yang Z.; Yuang X.-j.; Zhu R.-k.; Hou L.-h.; Jia J.-h.; Wang L.; Wu S.-t.; Bai B.; Gao X.-h.; Yang C. New Advance in Unconventional Petroleum Exploration and Research in China. Bull. Mineral., Petrol. Geochem. 2012, 31, 312–322. [Google Scholar]
- Tong X.; Zhang G.; Wang Z.; Tian Z.; Niu J.; Wen Z. Global oil and gas potential and distribution. Earth Sci. Front. 2014, 21, 1–9. [Google Scholar]
- Zou C. N.; Zhai G. M.; Zhang G. Y.; Wang H. J.; Zhang G. S.; Li J. Z.; Wang Z. M.; Wen Z. X.; Ma F.; Liang Y. B.; Yang Z.; Li X.; Liang K. Formation, distribution, potential and prediction of global conventional and unconventional hydrocarbon resources. Pet. Explor. Dev. 2015, 42, 14–28. 10.1016/S1876-3804(15)60002-7. [DOI] [Google Scholar]
- Liu P.; Wang X. F.; Lin Y.; Liu C. J.; Li X. F.; Liu W. H. Chemical and carbon isotope fractionations of alkane gases desorbed from confined systems and the application toward shale gas reservoir. Mar. Pet. Geol. 2020, 113, 104103 10.1016/j.marpetgeo.2019.104103. [DOI] [Google Scholar]
- Liu P.; Wang X. F.; Li X. F.; Zhang T.; Du G. C.; Liu W. H. Competitive adsorption characteristics of CH4/C2H6 gas mixtures on model substances, coal and shale. Fuel 2020, 279, 118038 10.1016/j.fuel.2020.118038. [DOI] [Google Scholar]
- Jia C. Z.; Zheng M.; Zhang Y. F. Unconventional hydrocarbon resources in China and the prospect of exploration and development. Pet. Explor. Dev. 2012, 39, 139–146. 10.1016/S1876-3804(12)60026-3. [DOI] [Google Scholar]
- Yao J.; Tang J.; Pang G.; Ma X.; Wang Q. Quantitative Simulation on Porosity-evolution in Member 8 of Yanchang Formation of Baibao-Huachi Area, Ordos Basin. Nat. Gas Geosci. 2013, 24, 38–46. [Google Scholar]
- Yang H.; Liu X.; Yan X. The relationship between tectonic-sedimentary evolution and tight sandstone gas reservoir since the late Paleozoic in Ordos Basin. Earth Sci. Front. 2015, 22, 174–183. [Google Scholar]
- Wang Y.; Liu L. F.; Li S. T.; Ji H. T.; Xu Z. J.; Luo Z. H.; Xu T.; Li L. Z. The forming mechanism and process of tight oil sand reservoirs: A case study of Chang 8 oil layers of the Upper Triassic Yanchang Formation in the western Jiyuan area of the Ordos Basin, China. J. Pet. Sci. Eng. 2017, 158, 29–46. 10.1016/j.petrol.2017.08.026. [DOI] [Google Scholar]
- Zhong D. Micro-petrology, pore throat characteristics and genetic mechanism of tight oil reservoirs-A case from the 6∼(th) and 7∼(th) members of Triassic Yanchang Formation in the Ordos Basin. Oil Gas Geol. 2017, 38, 49–61. [Google Scholar]
- Li S.; Wang Q.; Zhong J.; Qiu J.; Liu Z.; Qiao H.; Liao P. The Genesis of Sandbody in the Shallow Delta from Chang 8_1 of Jiyuan Area in Ordos Baisn. Nat. Gas Geosci. 2013, 24, 1102–1108. [Google Scholar]
- Cai Y.; Xiong Q.; Li Y.; Ding Y.; Xu W. Influence of pore structure on reservoir quality of low permeability sandstone reservoir: A case study from Chang 8 oil reservoir set in Jiyuan area, Ordos Basin. Lithol. Reservoirs 2014, 26, 69–74. [Google Scholar]
- Lan Y. F.; Deng X. Q.; Cheng D. X.; Huang J. X. Diagenetic facies and reservoir quality evaluation of Chang 6 sandstone reservoir in the upper Triassic Yanchang Formation of Huaqing area, Ordos Basin. Acta Petrol. Mineral. 2013, 33, 51–63. [Google Scholar]
- Li Y.; Chang X. C.; Yin W.; Sun T. T.; Song T. T. Quantitative impact of diagenesis on reservoir quality of the Triassic Chang 6 tight oil sandstones, Zhenjing area, Ordos Basin, China. Mar. Pet. Geol. 2017, 86, 1014–1028. 10.1016/j.marpetgeo.2017.07.005. [DOI] [Google Scholar]
- Xu L.; Niu X.; Liang X.; Chen Y.; Ma J. Microscopic Characteristics of Pore-throat Structures of the Chang 6 Member Tight Oil Layer in the Huaqing Region. J. Southwest Pet. Univ., Sci. Technol. Ed. 2017, 39, 36–46. 10.11885/j.issn.16745086.2016.03.28.03. [DOI] [Google Scholar]
- Houseknecht D. W. Assessing the relative importance of compaction processes and cementation to reduction of porosity in sandstones. AAPG Bull. 1987, 71, 633–642. 10.1306/9488787F-1704-11D7-8645000102C1865D. [DOI] [Google Scholar]
- Ehrenberg S. N. Measuring sandstone compaction from modal analyses of thin sections; how to do it and what the results mean. J. Sediment. Res. 1995, 65, 369–379. 10.1306/D42680C7-2B26-11D7-8648000102C1865D. [DOI] [Google Scholar]
- Damanti J. F.; Jordan T. E. Cementation and compaction history of synorogenic foreland basin sedimentary rocks from Huaco, Argentina. AAPG Bull. 1989, 73, 858–873. 10.1306/44B4A280-170A-11D7-8645000102C1865D. [DOI] [Google Scholar]
- Wilson J. C.; McBride E. F. Compaction and Porosity Evolution of Pliocene Sandstones, Ventura Basin, California: REPLY1. AAPG Bull. 1989, 73, 1280 10.1306/44B4AA28-170A-11D7-8645000102C1865D. [DOI] [Google Scholar]
- Dai J.; Zhang Y.; Xiong Q.; Wang Z. Effects of diagensis on reservoir property and auality, a case study of the Cainan Oilfield in east of Zhun Ga’er Basin. Pet. Explor. Dev. 2003, 30, 54. [Google Scholar]
- Jinglan L.; Xiaohong Li.; Tong L.; San Z.; Bo L. Impact of Diagenesis and Hydrocarbon Emplacement on Sandstone Reservoir Quality of the Yanchang Formation (Upper Triassic) in the Ordos Basin. Acta Geol. Sin. 2006, 80, 664–673. [Google Scholar]
- Meng Y.; Wang Y.; Jiang W.; Zhang J.; Liu D.; Niu J.; Sun H.; Shi S.; Gu T.; Xiu H. Porosity evolution simulation of the Member 4 of Shahejie Formation of Paleogene in Shuangqing area, Liaohe Depression. J. Palaeogeogr. 2009, 11, 225–232. [Google Scholar]
- Pan G.; Liu Z.; Zhao S.; Hu Z.; Hu X. Quantitative simulation of sandstone porosity evolution:A case from Yanchang Formation of the Zhenjing area, Ordos Basin. Acta Pet. Sin. 2011, 32, 249–256. [Google Scholar]
- Zhang S.; Ding X.; Wan Y.; Xiong D.; Zhu Z. Formation Mechanism and Distribution of Clay Minerals of Deeply Tight Siliciclastic Reservoirs. J. Southwest Pet. Univ. 2012, 34, 174–182. [Google Scholar]
- Yao J. L.; Deng X. Q.; Zhao Y. D.; Han T. Y.; Chu M. J.; Pang J. L. Characteristics of tight oil in Triassic Yanchang Formation, Ordos Basin. Pet. Explor. Dev. 2013, 40, 161–169. 10.1016/S1876-3804(13)60019-1. [DOI] [Google Scholar]
- Huang S.; Xie L.; Zhang M.; Wu W.; Shen L.; Liu J. Formation mechanism of authigenic chlorite and relation to preservation of porosity in nonmarine Triassic reservoir sandstones, Ordos Basin and Sichuan Basin, China. J. Chengdu Univ. Technol., Sci. Technol. Ed. 2004, 31, 273–281. [Google Scholar]
- Zhixue S.; Sun Z.; Lu H.; Yin X. Characteristics of carbonate cements in sandstone reservoirs:A case from Yanchang Formation, middle and southern Ordos Basin, China. Pet. Explor. Dev. 2010, 37, 543–551. 10.1016/S1876-3804(10)60054-7. [DOI] [Google Scholar]
- Zhao C. L.; Zhao B.; Shi Z. X.; Xiao L.; Wang D. H.; Khanchuk A. I.; Ivanov V. V.; Blokhin M. G. Maceral, mineralogical and geochemical characteristics of the Jurassic coals in Ningdong Coalfield, Ordos Basin. Energy Explor. Exploit. 2014, 32, 965–987. 10.1260/0144-5987.32.6.965. [DOI] [Google Scholar]
- Wang Y.; Liu L. F.; Cheng H. F. Gas Adsorption Characterization of Pore Structure of Organic-rich Shale: Insights into Contribution of Organic Matter to Shale Pore Network. Nat. Resour. Res. 2021, 30, 2377–2395. 10.1007/s11053-021-09817-5. [DOI] [Google Scholar]
- Zhu X.; Deng X.; Liu Z.; Sun B.; Liao J.; Hui X. Sedimentary characteristics and model of shallow braided delta in large-scale lacustrine: An example from Triassic Yanchang Formation in Ordos Basin. Earth Sci. Front. 2013, 20, 19–28. [Google Scholar]
- Guo R.; Zhao Y.; Wang W.; Hu X.; Zhou X.; Hao L.; Ma X.; Ma D.; Li S. Application of Rare-Earth Elements and Comparison to Molecular Markers in Oil–Source Correlation of Tight Oil: A Case Study of Chang 7 of the Upper Triassic Yanchang Formation in Longdong Area, Ordos Basin, China. ACS Omega 2020, 5, 22140–22156. 10.1021/acsomega.0c02233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang J.; Li S.; Zhou X.; Liang Y.; Guo R.; Yan C.; Chen J.; Li S. Development pattern and genesis of gentle slope shallow water delta front sand bodies of Chang 8_2 sand set in Zhidan area,Ordos Basin. Lithol. Reservoirs 2020, 32, 36–50. 10.12108/yxyqc.20200105. [DOI] [Google Scholar]
- Mou W.; Li S.; Yan C. Genesis Model of Sandbodies in Chang6 Tight Oil Reservoirs in the Western Area of Jiyuan, the Ordos Basin. Spec. Oil Gas Reservoirs 2017, 24, 6–11. [Google Scholar]
- Changyong W.; Rongcai Z.; Haihong W.; Yonglin H. A. N.; Chengyu W.; Xiaobing N. I. U. Provenance Analysis of Chang-6 Oil Reservoir Set in Jiyuan Region in Ordos Basin. Acta Sedimentol. Sin. 2008, 26, 933–938. [Google Scholar]
- Scherer M. Parameters Influencing Porosity in Sandstones: A Model for Sandstone Porosity Prediction1. AAPG Bull. 1987, 71, 485–491. 10.1306/703C80FB-1707-11D7-8645000102C1865D. [DOI] [Google Scholar]
- Lai J.; Wang G. W.; Ran Y.; Zhou Z. L.; Cui Y. F. Impact of diagenesis on the reservoir quality of tight oil sandstones: The case of Upper Triassic Yanchang Formation Chang 7 oil layers in Ordos Basin, China. J. Pet. Sci. Eng. 2016, 145, 54–65. 10.1016/j.petrol.2016.03.009. [DOI] [Google Scholar]
- Folk R. L.; Andrews P. B.; Lewis D. Detrital sedimentary rock classification and nomenclature for use in New Zealand. N. Z. J. Geol. Geophys. 1970, 13, 937–968. 10.1080/00288306.1970.10418211. [DOI] [Google Scholar]
- Kuang L. C.; Tang Y.; Lei D. W.; Chang Q. S.; Ouyang M.; Hou L. H.; Liu D. G. Formation conditions and exploration potential of tight oil in the Permian saline lacustrine dolomitic rock, Junggar Basin, NW China. Pet. Explor. Dev. 2012, 39, 700–711. 10.1016/S1876-3804(12)60095-0. [DOI] [Google Scholar]
- Guo R.; Xie Q.; Qu X.; Chu M.; Li S.; Ma D.; Ma X. Fractal characteristics of pore-throat structure and permeability estimation of tight sandstone reservoirs: A case study of Chang 7 of the Upper Triassic Yanchang Formation in Longdong area, Ordos Basin, China. J. Pet. Sci. Eng. 2020, 184, 106555 10.1016/j.petrol.2019.106555. [DOI] [Google Scholar]
- Wang Y.; Liu L. F.; Cheng H. F. Pore structure of Triassic Yanchang mudstone, Ordos Basin: Insights into the impact of solvent extraction on porosity in lacustrine mudstone within the oil window. J. Pet. Sci. Eng. 2020, 195, 107944 10.1016/j.petrol.2020.107944. [DOI] [Google Scholar]
- Zhao X.; Yang Z.; Lin W.; Xiong S.; Luo Y.; Wang Z.; Chen T.; Xia D.; Wu Z. Study on Pore Structures of Tight Sandstone Reservoirs Based on Nitrogen Adsorption, High-Pressure Mercury Intrusion, and Rate-Controlled Mercury Intrusion. J. Energy Resour. Technol. 2019, 141, 112903 10.1115/1.4043695. [DOI] [Google Scholar]
- Xiao Q.; Yang Z.; Wang Z.; Qi Z.; Wang X.; Xiong S. A full-scale characterization method and application for pore-throat radius distribution in tight oil reservoirs. J. Pet. Sci. Eng. 2020, 187, 106857 10.1016/j.petrol.2019.106857. [DOI] [Google Scholar]
- Bjorkum P. A. How important is pressure in causing dissolution of quartz in sandstones?. J. Sediment. Res. 1996, 66, 147–154. 10.1306/D42682DE-2B26-11D7-8648000102C1865D. [DOI] [Google Scholar]
- Ren D.; Sun W.; Tian H.; Liu D.; Zhang X.; An L. Research on the characteristic of diagenesis and porosity evolution of Chang 6 Reservoir in Jiyuan Oilfield, Ordos Basin. J. Northwest Univ., Nat. Sci. Ed. 2016, 46, 105–112. [Google Scholar]
- Shackleton N.; Kennett J. Paleotemperature History of the Cenozoic and the Initiation of Antarctic Glaciation: Oxygen and Carbon Isotope Analyses in DSDP Sites 277, 279 and 281. Initial Rep. Deep Sea Drill. Proj. 1975, 29, 743–756. 10.2973/DSDP.PROC.29.117.1975. [DOI] [Google Scholar]
- Walderhaug O. Temperatures of quartz cementation in Jurassic sandstones from the Norwegian continental shelf; evidence from fluid inclusions. J. Sediment. Res. 1994, 64, 311–323. 10.1306/D4267D89-2B26-11D7-8648000102C1865D. [DOI] [Google Scholar]
- Rezaee M. R.; Tingate P. R. Origin of quartz cement in the Tirrawarra Sandstone, southern Cooper Basin, South Australia. J. Sediment. Res. 1997, 67, 168–177. 10.1306/D4268522-2B26-11D7-8648000102C1865D. [DOI] [Google Scholar]













