You are on page 1of 15

Journal of Natural Gas Science and Engineering 76 (2020) 103189

Contents lists available at ScienceDirect

Journal of Natural Gas Science and Engineering


journal homepage: http://www.elsevier.com/locate/jngse

Experimental investigation of the petrophysical properties, minerals,


elements and pore structures in tight sandstones
Shuai Yin a, b, c, *, Li Dong d, **, Xia Yang e, Ruyue Wang d, ***
a
School of Earth Science and Engineering, Xi’an Shiyou University, Xi’an, 710065, China
b
Key Laboratory of Tectonics and Petroleum Resources (China University of Geosciences), Ministry of Education, Wuhan, 430074, China
c
Shanxi Key Laboratory of Petroleum Accumulation Geology, School of Earth Science and Engineering, Xi’an Shiyou University, Shanxi Xi’an, 710065, China
d
SINOPEC Petroleum Exploration and Production Research Institute, Beijing, 100083, China
e
Shandong Geo-Mineral Engineering Group Co. Ltd., Jinan, Shandong, 250200, China

A R T I C L E I N F O A B S T R A C T

Keywords: Geological conditions on the eastern margin of the Ordos Basin are complex, and the amount of oil and gas
Ordos basin exploration is low. In this paper, taking the S block in the eastern Ordos Basin as an example, we investigated the
Tight gas sandstone petrophysical properties, minerals, elements and pore structures of the middle Permian Shihezi Formation tight
Petrophysical properties
sandstones. The results show that the main lithologies of the target layer are lithic feldspar sandstone, lithic
Minerals
quartz sandstone, feldspar lithic sandstone and lithic sandstone. The pore types include residual intergranular
Elements
Pore structure pores, intergranular dissolution pores, intragranular dissolution pores, intercrystalline pores and microfractures.
The porosity of the rock samples ranges between 1.4% and 19.5%, with an average of 9.95%; the permeability
varies between 0.030 mD and 1.941 mD, with an average of 0.342 mD; and the water saturation is distributed
between 1.0% and 69.5%, with an average of 30.7%. The water saturation of rock samples has negative cor­
relations with porosity and permeability. Both displacement pressure (Pd) and median pressure (Pm) have good
negative correlations with rock permeability. The median pore throat radius (rm) has a good positive correlation
with rock permeability. Large pores with a throat radius greater than 0.1 μm are most favorable for rock
permeability. The Al/(Al þ Fe þ Mn) values of the rock samples range from 0.67 to 0.80, indicating that the silica
in the rock was not affected by hydrothermal fluid during the formation process. The Si/(Si þ Al þ Fe) values of
the rock samples range from 0.69 to 0.78, with an average of 0.73. Si has good positive correlations with Si/Al,
Al/(Al þ Fe þ Mn) and Si/(Si þ Al þ Fe). The sedimentary environment of the Shihezi Formation sandstone was
characterized by fresh water, a dry-hot climate and oxidizing conditions. The major elements correlate well with
the pore structure parameters: as the debris particles or quartz grains increase, the petrophysical properties
gradually improve. Finally, based on the petrophysical properties, minerals, elements and pore structure pa­
rameters, classification criteria for the pore structure of the Shihezi Formation tight sandstones in the study area
are proposed.

1. Introduction Formation and the He 8 segment of the middle Permian Shihezi For­
mation (Dou et al., 2010). However, the reservoir-forming conditions
The upper Paleozoic Carboniferous-Permian units in the Ordos Basin and natural gas enrichment patterns of tight sandstone reservoirs in the
of China are rich in tight sandstone gas resources, which are mainly marginal regions of the Ordos Basin are complex, and the current
distributed in the central part of the basin (Cao et al., 2019; Du et al., research level is very limited (Qiu et al., 2015a; Zhang et al., 2010).
2019; Fu et al., 2016; Han et al., 2019; Li et al., 2017; Lai et al., 2018a; Compared with the central part of the basin, the eastern margin of the
Yin et al., 2019a; Zou et al., 2012). Natural gas in the horizon is mainly Ordos Basin contains upper Paleozoic sand bodies that are more devel­
distributed in the Shan 1 segment of the lower Permian Shanxi oped, and natural gas is found from the lower Carboniferous Benxi

* Corresponding author. School of Earth Science and Engineering, Xi’an Shiyou University, Xi’an, 710065, China.
** Corresponding author.
*** Corresponding author.
E-mail addresses: speedysys@163.com (S. Yin), dongli.syky@sinopec.com (L. Dong), wry1990@vip.qq.com (R. Wang).

https://doi.org/10.1016/j.jngse.2020.103189
Received 16 September 2019; Received in revised form 30 January 2020; Accepted 31 January 2020
Available online 1 February 2020
1875-5100/© 2020 Elsevier B.V. All rights reserved.
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 1. Location and structure of the study area. The study area (S block) is located on the eastern edge of the Ordos Basin. The structural location of this area is
between the Yishan Slope and the Jinxi Flexural Belt. (a) Location of the study area (S block); (b) Contour map at the bottom of the Shihezi Formation.

Fig. 2. Stratigraphic unit division and sedimentary facies types in the upper Paleozoic strata of the study area. The target layer of this study is the middle Permian
Shihezi Formation.

Formation to the upper Permian Shiqianfeng Formation (Pang et al., degrees of natural gas enrichment for different sand bodies are very
2019; Qiu et al., 2014; Shi et al., 2019; Yin and Gao., 2019). Natural gas significant. The drilling of many low-productivity ineffective gas wells
is widely distributed in different layers, but the differences in the has severely increased both development costs and the risk of natural

2
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

and element experiments. This study can provide a basis for the evalu­
ation of tight sandstone gas reservoirs.

2. Geological background

The study area is located in the S block on the eastern edge of the
Ordos Basin (Fig. 1a). The structural location of this area is between the
Yishan Slope and the Jinxi Flexural Belt, which is a westward-dipping
monocline (Fu et al., 2016). The central and southern parts of the
study area have greater burial depths, and the strata are relatively flat,
while the northeastern part has a shallow depth and the formation dip is
relatively steep (Fig. 1b). This area was a relative uplifted region from
the middle and late Proterozoic to the Paleozoic, and the deposition rate
was low before the early Permian. Then, this area continued to rise after
the Late Jurassic and eventually became part of the westward-dipping
monocline of the Ordos Basin (Wang et al., 2017; Yin et al., 2019a).
The tectonic evolution of the Jinxi Flexural Belt is controlled by the
eastern Lvliangshan uplift and basement faults. The fold and fault ac­
tivity was mainly influenced by the Yanshanian movement and finally
shaped during the Himalayan movement. The structural traces of the
folds are mainly in the N–S direction. The strata have undergone strong
denudation since the Cretaceous, and the denudation thickness ranges
mainly between 1300 m and 2000 m (Fu et al., 2016; Yang et al., 2017).
Jurassic and Cretaceous strata in the study area are missing.
The stratigraphic unit divisions and sedimentary facies descriptions
Fig. 3. Lithologic triangular diagram of the Shihezi Formation sandstones in
of the upper Paleozoic rocks in the study area are shown in Fig. 2. The
the study area. The results indicate that the Shihezi Formation sandstones are
target layer of this study is the middle Permian Shihezi Formation. The
mainly lithic feldspathic sandstone, lithic quartz sandstone, feldspar lithic
sandstone and lithic sandstone.
lower part of the Shihezi Formation (segments He 5 - He 8) consists of a
braided delta plain facies, and the upper part (segments He 1 - He 4)
contains a meandering river delta plain facies. The thickness of the
gas exploration (Dutton and Loucks., 2010; Morad et al., 2010; Nelson,
sandstone at the bottom of the Shihezi Formation is relatively large, and
2009). Moreover, some wells have serious water production, which
the thickness of a single sand body can reach 15 m. The thickness of the
seriously affects the economics of tight sandstone gas development.
sand bodies gradually becomes thinner upward, and the thickness of a
The upper Paleozoic tight gas sandstones in the Ordos Basin have
single sand body in the upper part of the formation is usually less than
poor petrophysical properties, a low formation pressure coefficient and
10 m.
insufficient natural energy, and the pore structure is an important factor
that affects and determines the fluid flow and hydrocarbon accumula­
3. Methodology
tion in the reservoirs (Lai and Wang., 2015; Rezaee et al., 2012; Xi et al.,
2016). After years of exploration, petroleum geology researchers have
In this paper, we obtained thin section observations, scanning elec­
discovered that sedimentation is an important factor controlling the
tron microscopy images, and petrophysical properties and conducted X-
formation of upper Paleozoic gas reservoirs in this area and that reser­
ray diffraction, high-pressure mercury intrusion and mineral and
voir properties are the key factors determining the distribution of gas
element experimental tests on tight sandstone samples of the Shihezi
reservoirs (Fu et al., 2016; Yang et al., 2017). Previous studies on tight
Formation from the study area.
gas reservoirs in this area have mainly focused on seismic and logging
identification of sedimentary facies, sedimentary microfacies charac­
3.1. Thin section and scanning electron microscope (SEM) imaging
teristics and distribution patterns, and reservoir anisotropy (Fu et al.,
2016; Jiang et al., 2016; Zou et al., 2012). However, relatively few
Thin section and SEM observations were performed on 45 groups of
studies have focused on the characteristics of the reservoir’s micropore
sandstone samples from the target layer. The mineral compositions and
structures. Therefore, a systematic study of the microscopic pore struc­
optical properties of the thin sections were determined with a polarizing
ture of rocks can lead to a more accurate understanding of the real sit­
microscope. The device was a Leica DM4500 microscope. The thin
uation of underground reservoirs (Anovitz and Cole., 2015; Chang et al.,
section size was 25 mm � 25 mm, and the thickness was 0.03 mm. For
2019; Liu et al., 2017a; Shanley and Cluff., 2015; Wang et al., 2017; Yin
SEM observations, the equipment was an FEI-200F system. A fine-
and Ding., 2019). In addition, the study of the microscopic pore struc­
focusing electron beam was used to perform a point-by-point scan on
ture of tight sandstones is highly significant for the design and adjust­
the surface of the sample to excite various electronic signals with
ment of well patterns in the later development stage (Abbaszadeh et al.,
different functions, and the detector selectively collected and processed
1996; Li et al., 2017; Xiao et al., 2016; Yin et al., 2019b, 2019c).
the desired electronic signals and converted the signals into images. The
Geological conditions in the eastern Ordos Basin are complex, and
test standard was SY/T 5162–2014.
the amount of oil and gas exploration is low. Strengthening the study of
microscopic pore structure characteristics of tight gas sandstones is
3.2. Porosity, permeability and water saturation
conducive to accelerating the exploration and development of lithologic
reservoirs in this area. In this paper, taking the S block in the eastern
Porosity and permeability tests were performed on 40 groups of
Ordos Basin as an example, a systematic study was conducted to
sandstone samples from the target layer. The test instruments for
investigate the petrophysical properties, minerals, elements and pore
porosity and permeability were HKC-2 and HBSST-3, respectively. The
structures of the middle Permian Shihezi Formation tight sandstones
sample size was a small column with a diameter of 25 mm and a length
using thin sections, scanning electron microscopy, petrophysical prop­
of 50 mm. Porosity was determined by the volumetric method, and the
erty tests, X-ray diffraction, high-pressure mercury injection and mineral
test standard was SY/T 5336–2006. The measured rock permeability

3
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 4. Photomicrographs of the tight sandstone samples from the Shihezi Formation. On the left is an image of the sample under plane-polarized light, and on the
right is an image of the same sample under cross-polarized light. Notes: (a) Well G2, 1572.77 m, � 40; (b) Well G2, 1692.50 m, � 40; (c) Well G3-1, 1631.86 m, � 40;
(d) Well G5, 1503.53 m, � 40. Q-quartz; R-rock debris; F-feldspar.

4
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 5. Mineral compositions and contents of the tight sandstone samples.

Table 1
Petrophysical properties and mercury intrusion parameters of the tight sandstone samples in the target layer.
3
Well name No. H (m) Р (g⋅cm ) φ (%) K (mD) Sw (%) Mercury intrusion parameters

Pd (MPa) Pm (MPa) Smax (%) Sr (%) rmax (μm) rm (μm) DM So

HF-T36 S1 1651.3 2.65 8.2 0.322 / 0.24 5.00 89.18 37.89 3.086 0.147 11.74 2.59
HF-T36 S2 1689.5 2.69 13.2 0.797 / 0.24 4.89 89.49 27.83 3.108 0.150 11.15 2.08
HF-T36 S3 1714.2 2.64 12.5 0.224 / 0.71 7.18 94.52 40.08 1.030 0.102 11.90 1.88
HF-T38 S4 1659.3 2.66 12.2 0.342 44.8 0.24 8.36 88.33 35.67 3.127 0.088 11.80 1.93
HF-T38 S5 1667.2 2.74 11.0 0.091 / 0.71 14.55 88.62 42.30 1.033 0.051 12.55 1.86
HF-T38 S6 1680.9 2.66 11.4 0.448 / 0.34 7.32 92.77 30.05 2.145 0.100 11.67 1.91
HF-T41 S7 1626.8 2.68 10.5 0.153 69.5 0.50 16.03 93.72 47.57 1.476 0.046 12.57 2.03
HF-T41 S8 1652.0 2.67 15.3 0.252 / 0.72 11.69 88.56 41.66 1.026 0.063 12.24 1.74
HF-T41 S9 1686.3 2.65 11.6 0.410 / 1.02 16.64 90.44 42.25 0.720 0.044 12.98 1.81
G3-1 S10 1629.6 2.69 11.9 0.180 51.0 0.72 12.05 88.99 33.05 1.025 0.061 12.41 1.97
G3-1 S11 1631.2 2.70 7.5 0.050 40.0 1.46 30.23 86.13 30.01 0.504 0.024 13.52 1.77
G3-1 S12 1633.2 2.70 13.8 0.061 27.2 1.03 10.10 89.11 34.29 0.714 0.073 12.35 1.66
G3-1 S13 1634.5 2.69 9.0 0.068 51.6 1.03 14.89 91.37 36.32 0.716 0.049 12.88 1.55
G7 S14 1656.9 / 8.1 0.120 32.3 0.72 7.61 91.18 35.62 1.028 0.097 12.25 1.83
G7 S15 1657.9 / 8.8 0.207 33.9 0.72 6.26 87.08 41.31 1.028 0.118 11.92 1.86
G7 S16 1661.6 / 5.6 0.030 / 1.03 8.41 84.26 25.03 0.716 0.087 12.57 1.63
G7 S17 1693.4 2.69 12.7 0.443 46.8 0.35 24.39 89.75 49.91 2.129 0.030 12.86 2.07
G7 S18 1694.7 / 5.5 0.039 / 1.02 63.12 74.84 32.36 0.719 0.012 13.85 1.96
G7 S19 1696.1 2.71 4.9 0.042 / 1.02 42.89 72.63 40.48 0.718 0.017 12.86 2.24
G7 S20 1698.4 2.69 7.3 0.059 / 1.02 31.84 88.25 34.51 0.718 0.023 13.59 1.70
G7 S21 1699.8 / 7.3 0.111 / 1.02 13.13 88.35 39.70 0.718 0.056 12.66 1.60
G8 S22 1629.6 2.67 14.0 0.840 37.6 0.34 8.26 88.54 52.00 2.136 0.089 11.61 2.04
G8 S23 1632.3 2.65 12.2 0.830 41.1 0.35 4.46 89.86 50.55 2.123 0.165 11.31 2.07
G8 S24 1633.2 2.69 10.0 0.320 / 0.72 7.39 87.19 37.78 1.024 0.100 11.86 2.04
G9 S25 1629.6 / 10.0 0.560 25.9 0.50 3.39 90.25 32.52 1.477 0.217 11.08 1.78
G9 S26 1630.7 / 9.1 0.500 / 0.34 8.87 88.32 40.72 2.137 0.083 11.86 2.32
G9 S27 1633.3 2.66 6.5 0.052 40.5 0.49 13.98 86.40 25.86 1.493 0.053 13.13 1.59
G9 S28 1634.6 2.66 10.0 0.380 / 0.71 7.60 92.38 42.02 1.032 0.097 11.82 1.90
G11 S29 1643.1 2.68 8.3 0.186 / 0.72 12.85 89.93 43.86 1.028 0.057 12.40 2.00
G11 S30 1660.7 2.69 5.8 0.120 / 0.71 9.75 86.00 43.06 1.033 0.075 12.26 2.30
G11 S31 1672.4 2.66 10.8 0.556 / 0.35 7.34 89.32 35.25 2.124 0.100 11.68 2.19
G17 S32 1716.3 2.70 1.4 0.111 62.4 0.34 38.44 65.38 32.85 2.142 0.019 12.36 2.41
G17 S33 1725.9 2.70 5.1 0.156 14.6 0.50 61.22 72.25 32.25 1.478 0.012 13.65 2.07
G17 S34 1731.8 2.68 9.0 0.211 9.1 0.49 27.81 88.49 37.28 1.487 0.026 13.07 2.07
G17 S35 1739.2 2.69 8.5 0.210 17.6 0.50 17.26 92.70 38.24 1.469 0.043 12.73 1.98
G19 S36 1685.2 2.69 7.3 0.114 18.5 0.34 4.04 87.97 44.46 2.150 0.182 11.22 1.88
G19 S37 1687.1 2.68 11.4 0.664 1.7 0.34 2.17 90.24 47.48 2.158 0.339 10.60 1.66
G21 S38 1618.0 2.69 19.5 1.219 2.9 0.23 3.52 91.35 45.31 3.134 0.209 11.11 1.95
G21 S39 1620.6 2.67 19.2 1.941 1.0 0.24 2.07 95.22 34.43 3.113 0.357 10.50 1.82
G21 S40 1634.8 2.70 11.6 0.254 5.1 0.50 4.93 91.40 42.86 1.486 0.149 11.50 1.76

Notes: H – sampling depth; ρ – rock density; φ – rock porosity; k – rock permeability; Sw – water saturation; Pd – displacement pressure; Pm – median pressure; Smax –
maximum mercury saturation; Sr – mercury removal efficiency; rmax – maximum pore throat radius; rm – median pore throat radius; DM – throat means coefficient; So –
sorting factor.

5
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

groups of sandstone samples from the target layer. The test standard was
GB/T 29171–2012.

3.5. Major and trace elements

The major elements were measured with an Axios test instrument.


The element contents were determined based on the X-ray fluorescence
intensities of the elements. The test temperature was 23 � C, the relative
humidity was 40%, and the test standard was GB/T 21114–2007. For
trace elements, the test instrument was an inductively coupled plasma
mass spectrometer 300�. Major and trace elements were determined on
11 groups of sandstone samples from the target layer. The test method
was the ICP-MS external standard method. The test standard was GB/T
21114–2007.

4. Results

4.1. Lithologies and mineral compositions


Fig. 6. Relationship between rock porosity and permeability of the tight
sandstone samples. The thin section identifications indicate that the Shihezi Formation
sandstones are mainly lithic feldspathic sandstone, lithic quartz sand­
was the gas permeability, and the test standard was SY/T 5336–2006. stone, feldspar lithic sandstone and lithic sandstone (Figs. 3 and 4). The
Water saturation tests were performed on 22 groups of sandstone sam­ quartz particles in the rock are relatively clean, the particle size gener­
ples from the target layer. The water saturation was measured using a ally ranges between 0.25 mm and 0.50 mm, and quartz overgrowths can
Hongbo YBG-1 instrument, and the test standard was SYT 5336–2006. be observed. Feldspars include potassium feldspar, plagioclase, micro-
plagioclase, striped feldspar, etc., and polysynthetic twins and cross-
3.3. X-ray diffraction analysis hatched twins are visible. The particle size is generally distributed be­
tween 0.25 mm and 0.50 mm. Some feldspars have different degrees of
Each mineral crystal has a specific X-ray diffraction pattern, and the alterations, such as argillization and sericitization. The rock debris
characteristic peak intensity in the spectrum is positively correlated with components are mainly metamorphic rock debris and igneous rock
the mineral content. The mineral content was determined using the “K debris. Among them, the contents of igneous rock debris and meta­
value method”. The X-ray diffractometer was a Rigaku SmartLab9 in­ morphic rock debris are similar, but the metamorphic rock debris con­
strument. X-ray diffraction analysis was performed on 40 groups of tent is slightly higher than that of igneous rock debris.
sandstone samples from the target layer. The test standard was SY/T X-ray diffraction analysis was used to analyze the total rock mineral
5163–2010. composition of the 38 groups of samples (Fig. 5). The results show that
the main mineral types in the tight sandstones of the Shihezi Formation
3.4. High-pressure mercury injection experiment include quartz, potash feldspar, plagioclase, calcite and clay and small
amounts of dolomite, siderite, hematite, halite and ankerite. The con­
Based on the pore volume percentage of the injected mercury and the tents of quartz range from 31.8 to 66.6%, with an average of 47.0%; the
corresponding pressure, the relationship between the capillary pressure contents of feldspar (potash feldspar þ plagioclase) are distributed be­
and the mercury saturation of the rock sample can be obtained. Since the tween 2.6% and 43.5%, with an average of 20.5%; the contents of calcite
surface tension and wet contact angle of mercury are relatively stable, vary from 0 to 19.3%, with an average of 3.6%; and the contents of clay
the capillary pressure curve measured by an injection-type mercury are distributed between 14.1% and 42.4%, with an average of 28.0%.
intrusion meter was used to obtain the pore size and distribution. The
experimental equipment was a PoreMaster 60 GT porosimeter with a 4.2. Petrophysical characteristics
test pressure range of 0–60 000 psia and an accuracy of �0.11% fso.
High-pressure mercury injection experiments were performed on 40 The porosity and permeability of 40 groups of sandstones were

Fig. 7. Relationships between water saturation and rock porosity and permeability of the tight sandstone samples.

6
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 8. Thin section and scanning electron microscopy images of the Shihezi Formation sandstone samples. Notes: (a) 1588.88 m, Well SM-22, coarse feldspar lithic
sandstone with quartz overgrowth; (b) 1618.00 m, Well G21, lithic sandstone with residual primary intergranular pore; (c) 1632.28 m, Well G8, feldspar lithic
sandstone with intergranular authigenic quartz and dissolution pits visible on quartz surface; (d) 2027.1 m, Well SM-4, feldspar lithic sandstone with feldspar particle
dissolution pore; (e) 1587.54 m, Well SM-22, feldspar lithic sandstone with dissolved skeletal potassium feldspar; (f) 1572.77 m, Well G2, feldspar lithic sandstone
with skeletal feldspar and dissolution pores; (g) 1672.2 m, Well SM-7, coarse-to medium-grained feldspar lithic sandstone with intergranular dissolution pore; (h)
1588.88 m, Well SM-22, feldspar lithic sandstone with honeycomb mixed layer illite/smectite and visible intercrystalline pores; (i) 1676.16 m, Well SM-9, medium-
coarse lithic sandstone with a microfracture cutting through multiple particles.

determined. The test results are shown in Table 1. The porosity of the significant negative correlation between these two parameters. This
rock samples is distributed between 1.4% and 19.5%, with an average conclusion is consistent with the results of this paper. The relatively low
value of 9.95%; the permeability is distributed between 0.030 mD and correlation of the test results is because the target layer contains mul­
1.941 mD, with an average value of 0.342 mD. There is a good positive tiple sandstone lithologies. The pore structure and water saturation of
correlation between the porosity and permeability of the samples different sandstones vary greatly. Different sandstone lithologies cause
(Fig. 6). The rock porosity is generally less than 15%, while the rock the dispersion in the overall data, but the fitted results can still represent
permeability is generally less than 1.5 mD. Under formation conditions, the correlation between water saturation and petrophysical parameters.
rocks usually contain a certain amount of water (Lasaga, 1984; Liu et al.,
2017b). In this case, water is an important component of the tight
sandstones, and cation exchange between water and rock can affect 4.3. Pore type
reservoir properties. Formation water actually represents a chemical
environment (Bloch et al., 2002; Haneef et al., 1993). The tight sand­ According to the results of thin section observations and scanning
stones contain many micropores, and the water is mainly distributed on electron microscopy, the pore types of the Shihezi Formation tight
the surfaces of these micropores (Ghanizadeh et al., 2015). Water sandstones can be divided into five categories: residual intergranular
saturation has a serious impact on reservoir properties and gas well pores, intergranular dissolution pores, intragranular dissolution pores,
productivity (Yin et al., 2018a). intercrystalline pores and microfractures. Intergranular dissolution
An analysis of the relationship between water saturation and rock pores and intragranular dissolution pores are dominant, followed by
petrophysical parameters shows that water saturation is negatively residual intergranular pores, intercrystalline pores and microfractures.
correlated with rock porosity and permeability (Fig. 7). The water The compaction of the studied sandstones is very strong, with
saturation of the samples is distributed between 1.0% and 69.5%, with concavo-convex contacts and line-concavo-convex contacts between the
an average value of 30.7%. Low porosity and low permeability sand­ particles, followed by point-suture line contacts (Fig. 8a). At the same
stones have higher water saturation. This result occurs because the low- time, the Shihezi Formation sandstone is strongly cemented, which is
porosity and low-permeability sandstones are denser and their pore size mainly characterized by quartz overgrowths (Fig. 8a). Although the
is smaller, so more moisture is present in the interior of the micropores cementation by quartz reduces the porosity of the rock, it resists the
to become bound water or free water. Zeng et al. (2005) analyzed the compaction of the rock and thus maintains permeability (Loucks et al.,
relationship between water saturation and permeability of tight sand­ 2009). The jagged edges of quartz particles are easily dissolved by fluid,
stone through systematic experimental tests. The results showed a while the smooth edges of the quartz particles are generally not
conducive to dissolution.

7
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 9. Relationships between mercury intrusion parameters and petrophysical properties of the tight sandstone samples.

reactions of iron oxide cladding (Abdullah and Eswaran., 2014). The


Table 2
intragranular dissolution pores are mainly developed in feldspar parti­
Major and trace element test results for the tight sandstone samples.
cles. The feldspar grains are selectively eroded along the cleavage planes
No. Depth Major element combination Trace element to form reticular and fluted pores (Fig. 8d–f). The long axes of the pores
(m) combination
are substantially parallel to the cleavage planes. Some of the feldspar
Si Si/ Si/(Si Al/(Al Sr/ Sr/ Ni/ particles are completely dissolved into mold pores (Fig. 8d). The intra­
(%) Al þ Al þ þ Fe þ Ba Cu Co
granular dissolution pores are affected by lithology. In sandstones with
Fe) Mn)
high feldspar contents, the intragranular micropores are more devel­
S1 1651.3 32.43 4.05 0.76 0.78 0.23 13.71 1.34 oped. Intergranular dissolution pores are formed by the selective erosion
S2 1689.5 32.09 4.38 0.78 0.80 0.29 8.55 2.21
S3 1714.2 32.13 3.76 0.74 0.76 0.20 32.64 1.44
of carbonate cements and the clay matrix (Wang et al., 2019). The
S4 1659.3 30.39 4.19 0.75 0.71 0.27 13.99 1.74 intergranular dissolution pores are mainly distributed between adjacent
S5 1667.2 29.99 3.26 0.70 0.72 0.25 12.51 1.70 crumb particles (Fig. 8g).
S6 1680.9 31.45 3.71 0.73 0.74 0.26 15.73 1.66 Clay mineral intercrystalline pores are also developed in the Shihezi
S7 1626.8 29.7 3.37 0.69 0.67 0.16 11.68 1.49
Formation sandstones (Fig. 8h). Clay mineral intercrystalline pores are
S8 1652.0 29.77 3.36 0.69 0.67 0.19 12.75 1.57
S29 1643.1 28.31 3.48 0.70 0.67 0.23 11.52 2.01 typically distributed in the pores between the eroded particles and the
S30 1660.7 29.20 3.37 0.70 0.69 0.29 6.55 1.14 rock debris particles. The target layer is in the middle-late diagenesis
S31 1672.4 29.49 3.69 0.74 0.75 0.26 17.07 1.44 stage, and the main clay minerals developed are illite, illite/smectite
mixed layers, kaolinite and chlorite. Due to the crystal growth of clay
minerals, intercrystalline pores usually have good orientation. The
A small number of primary intergranular pores are also developed in
intercrystalline pores have small radii, usually less than 5 μm, so their
the sandstones (Fig. 8b). These pores are mostly filled at the later sedi­
effect on the petrophysical properties of the rock is very limited. How­
mentation stage. The interstitial materials are mainly clay matrix and
ever, the presence of a clay cladding can delay cementation and thus
secondary quartz and iron particles. The primary intergranular pores are
maintain porosity.
mostly angular pores or irregular polygonal pores, and the pore radii are
In addition, microfractures are developed in the Shihezi Formation
relatively large, generally distributed between 50 μm and 100 μm.
sandstones (Fig. 8i). Microfractures have limited pore-enhancing effects
The Shihezi Formation sandstones are strongly dissolved and have
but can improve reservoir connectivity and provide effective channels
formed many intragranular and intergranular dissolution pores
for fluid migration (Nabawy et al., 2009; Yin et al., 2018b, 2018c).
(Fig. 8c–g). The authigenic quartz particles show significant dissolution
(dissolution pits) (Fig. 8c). A dissolution pit is a circular pit formed by
chemical dissolution, and its formation is mostly related to the redox

8
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 10. Relationship between major elements of the tight sandstone samples.

4.4. Pore structure parameter properties. Fig. 9a shows that there is a good negative correlation be­
tween displacement pressure and permeability.
We conducted high-pressure mercury intrusion tests on 40 groups of The median pressure (Pm) is the capillary pressure corresponding to a
samples. The test results for each mercury intrusion parameter are mercury content of 50%. The larger the Pm value is, the denser the rock
shown in Table 1. The maximum mercury saturation (Smax) of the and the smaller the production capacity. The smaller the Pm value is, the
samples is large, ranging from 65.38% to 95.22%, with an average of better the permeability of the rock and the higher the production ca­
87.77%. The displacement pressure (Pd) ranges from 0.23 to 1.46 MPa, pacity (Torabi et al., 2013). Therefore, Pm also has a good negative
with an average of 0.61 MPa. The median pressure (Pm) varies between correlation with rock permeability (Fig. 9b).
2.07 and 63.12 MPa, with an average of 15.05 MPa. The median pore The median pore throat radius (rm) is the radius of the throats when
throat radius (rm) ranges from 0.012 to 0.357 μm, with an average of Pm is 50%. This parameter can reflect the size of the total pore throats. It
0.095 μm. The throat mean coefficient (DM) is distributed between 10.50 can be seen from Fig. 9c that rm has a good positive correlation with rock
and 13.85, with an average of 12.20. The sorting factor (So) varies be­ permeability. As rm increases from 0.05 μm to 0.35 μm, the rock
tween 1.549 and 2.590, with an average of 1.937. permeability increases significantly from approximately 0.2 mD to 1.3
mD. Large pores are more favorable than small pores for rock
5. Discussion permeability.
The throat mean coefficient (DM) represents the average position of
According to the experimental results, the relationships among the the full pore throat distribution. The larger the DM value is, the more the
petrophysical properties, mineral and element compositions and pore rock capillary pressure curve tends to thin skewness. If narrow throats
structures of the Shihezi Formation tight sandstones were systematically dominate the entire pore throat system, reservoir properties or fluid
studied. seepage are less favorable. The DM of the samples has a good negative
correlation with the permeability (Fig. 9d). This result shows that as DM
5.1. Relationships between pore structure parameters and petrophysical increases, the petrophysical properties of the rock gradually deteriorate.
properties
5.2. Relationships among pore structure parameters, minerals and
Compared to rock porosity, rock permeability can better reflect gas elements
well productivity and petrophysical properties (Sakhaee and Bryant.,
2014; Yin et al., 2018d). Therefore, we analyze the relationships be­ Geochemical elements are closely related to sedimentary character­
tween rock permeability and pore structure parameters (Fig. 9). The istics. The major and trace elements of the 11 groups of samples were
displacement pressure (Pd) is the minimum pressure required for a wet measured. The test results are shown in Table 2. Si is usually derived
phase fluid to be displaced by a nonwetting phase fluid (Pittman, 1992). from terrigenous debris, which indicates the presence of quartz and
The lower the displacement pressure is, the better the reservoir other detrital minerals. Si has a positive correlation with quartz content

9
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 11. Relationships between major elements and pore structure parameters in the tight sandstone samples.

(Fig. 10a). The aggregation of Al is associated with terrigenous debris with the X-ray diffraction mineral component analysis (Fig. 10c).
and is commonly found in feldspar and clay minerals. It can be seen from Al/(Al þ Fe þ Mn) is an important index for discriminating the silica
Fig. 10b that Si has a good positive correlation with Si/Al. This result formation environment, and it has a good positive correlation with the Si
indicates that as the quartz content increases, the content of feldspar content (Fig. 10d). Al is related to terrigenous material, while Fe and Mn
plus clay minerals gradually decreases. This observation is consistent are related to hydrothermal fluid. An Al/(Al þ Fe þ Mn) ratio of 0.01

10
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Fig. 12. Mercury intrusion curves and pore throat radius distribution for different types of pore structure. The pore structures of the tight sandstones in the Shihezi
Formation are divided into Type I, Type II and Type III.

represents a pure hydrothermal fluid origin (Harris et al., 2011; Liu the depositional environment experienced oxidizing conditions
et al., 2018; Qiu et al., 2015b). The Al/(Al þ Fe þ Mn) values of the (Table 2).
Shihezi Formation sandstones in the study area are distributed between The relationships among minerals, elements and pore structure pa­
0.67 and 0.80, with an average of 0.72. These values indicate that the rameters were further compared. Si has good correlations with the
silica in the rock is not affected by hydrothermal fluid during the for­ median pressure, the median throat radius and the throat mean coeffi­
mation process and that there is some biogenic silica in the rock. There is cient (Fig. 11a–c). With an increase in detrital particles or quartz grains,
also a good positive correlation between Si/(Si þ Al þ Fe) and Si content the median pressure gradually decreases, the median throat radius in­
(Fig. 10d). The Si/(Si þ Al þ Fe) values of the Shihezi Formation creases, and the throat mean coefficient decreases. These relationships
sandstones in the study area vary between 0.69 and 0.78, with an represent a process in which the petrophysical properties of rocks
average of 0.73. improve. Production practices have also shown that gas production ca­
Trace elements are an important basis for the detailed study of pacity increases with increasing quartz content (Li et al., 2015). Fig. 11d
sedimentary depositional environments. The Sr/Ba values of the sam­ also shows a good negative correlation between Si/Al and the throat
ples range between 0.16 and 0.29, with an average of 0.24. Sr/Ba values mean coefficient. Both element combinations Al/(Al þ Fe þ Mn) and
less than 1 indicate that the depositional environment was fresh water. Si/(Si þ Al þ Fe) have good correlations with pore structure parameters
The Sr/Cu values vary between 6.55 and 32.64, with an average of (Fig. 11e–h). These results show that with increasing Al/(Al þ Fe þ Mn)
14.25. Sr/Cu values greater than 10 indicate that the depositional or Si/(Si þ Al þ Fe), the petrophysical properties of rock gradually
environment had a dry-hot climate. The Ni/Co values range from 1.14 to become better.
2.21, with an average of 1.61. Ni/Co values less than 2.5 indicate that

11
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

with an average Smax value of 87.77%, and nearly 90% of the samples
have Smax values greater than 80%. Therefore, these sandstones with
Smax greater than 80% should be finely classified. Referring to the pre­
vious classification criteria for the pore structure of tight sandstones, the
pore structures of the tight sandstones in the Shihezi Formation of the
study area are divided into Type I, Type II and Type III (Fig. 12). The
Smax values of Type I and Type II sandstones are greater than 80%, while
the Smax values of Type III sandstones are less than 80%. The charac­
teristics of these three types of pore structures are as follows.
Type I pore structure corresponds to high-quality reservoirs; the rock
porosity is generally greater than 10%, the rock permeability is greater
than 0.5 mD, and Pd and Pm are both small (Pd less than 0.5 MPa and Pm
less than 10 MPa). The pore types are dominated by intragranular and
intergranular dissolution pores, and some microfractures are also
developed. The pore diameter is generally greater than 100 μm, rm
ranges from 0.08 to 0.36 μm, and So is generally greater than 1.65. The
pore throat radius distribution of this type of reservoir features a single-
peak coarse throat (Fig. 12a); that is, the main peak is located on the side
with a larger pore throat radius. Large pore throats make the greatest
contribution to rock permeability.
Fig. 13. Cumulative porosity and permeability distribution curves of the S18 Type II pore structures correspond to a general reservoir; the rock
sandstone sample. porosity is mainly distributed from 6 to 12%, the rock permeability
ranges from 0.03 to 0.5 mD, Pd varies from 0.24 to 1.46 MPa, and Pm is
5.3. Pore structure classification usually less than 30 MPa. The pore type is dominated by intragranular
dissolution pores. The pore diameter is usually larger than 50 μm, rm is
According to the high-pressure mercury test results, porosity, distributed between 0.02 and 0.18 μm, and So is generally greater than
permeability and microscopic observations, the pore structure of the 1.55. The pore throat radius distribution of this type of reservoir is
tight sandstones in the Shihezi Formation was classified. A sandstone bimodal (Fig. 12b). That is, there are two main peak regions of large
with an Smax value greater than 80%, a porosity greater than 10–12%, a pores and small pores. However, the cumulative frequency of the main
permeability greater than 0.4–0.5 mD, and a single-peak coarse throat peak region of the large pores is larger than that of the small pores.
radius is usually defined as a type I pore structure in previous studies Type III pore structures correspond to poor reservoirs; the rock
(Ghanizadeh et al., 2015; Lai et al., 2018b; Nelson., 2009). The porosity is usually less than 6%, the rock permeability is usually less
maximum mercury saturations of the studied tight sandstones are large, than 0.15 mD, Pd is distributed in 0.34–1.02 MPa, and Pm is usually

Fig. 14. Capillary pressure curves of partial sandstone samples from the Shihezi Formation. (a) Correspondence between mercury saturation and capillary pressure;
(b) Pore structures of the samples are classified into Type I, Type II and Type III.

12
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Table 3
Pore structure classification standard for the tight sandstone reservoirs in the target layer.
Classification index Type I Type II Type III

Petrophysical φ >10% 6–12% <6%


parameter k >0.5 mD 0.03–0.5 mD <0.15 mD
Sw <40% 5–50% >50%
Pore structure Pore type Intragranular and intergranular Intragranular dissolution pores Few intragranular dissolution pores
parameter dissolution pores and few microcracks and intercrystalline pores
Pd <0.5 MPa 0.24–1.46 MPa 0.34–1.02 MPa
Pm <10 MPa <30 MPa >38 MPa
Smax >80% >80% <80%
Sr 28–89% 25–50% 30–40%
Pore diameter >100 μm >50 μm <50 μm
rm 0.08–0.36 μm 0.02–0.18 μm <0.02 μm
So >1.65 >1.55 1.96–2.41
DM <12.0 11.0–13.5 >12.4
Mercury intrusion curve Bias to the lower left Between the mercury injection curves of Bias to the upper right
morphology type I and type II pore structures
Pore throat radius Single-peak coarse throat type Double-peak coarse throat type Double-peak fine throat type
distribution
Element ratio Si/Al >3.65 3.30–4.20 <3.30
Si/(Si þ Al þ Fe) >0.73 0.70–0.73 <0.70
Al/(Al þ Fe þ Mn) >0.75 0.67–0.78 <0.67

Notes: φ – rock porosity; k – rock permeability; Sw – water saturation; Pd – displacement pressure; Pm – median pressure; Smax – maximum mercury saturation; Sr –
mercury removal efficiency; rm – median pore throat radius; DM – throat means coefficient; So – sorting factor.

greater than 38 MPa. The pore types include small numbers of intra­ lower left (Fig. 14b).
granular dissolution pores and intercrystalline pores, the pore diameter
is generally less than 50 μm, rm is less than 0.02 μm, and So is between 5.4. Pore structure classification criteria
1.96 and 2.41. The pore throat radius distribution of this type of reser­
voir is also bimodal (Fig. 12c). There are two main peak regions of large Based on the comprehensive study of petrophysical properties,
pores and small pores. However, the cumulative frequency of the main minerals, elements and pore structure parameters of the tight sandstone
peak region of the large pores is smaller than that of the small pores. Fine samples, classification criteria for the pore structure of tight sandstone
pores are mainly developed in this type of reservoir, and the rock in the Shihezi Formation are proposed (Table 3). Type I has the best
permeability is very low. reservoir properties, followed by Type II, while Type III indicates an
For the tight sandstone reservoirs of the Shihezi Formation in the ineffective reservoir. Therefore, tight sandstones with Type I pore
study area, the large pores have the most important effect on the rock structure should be the focus of sweet spot exploration for natural gas in
permeability. The cumulative porosity and permeability curves of the the Shihezi Formation.
tight sandstone samples also show that pores with a throat radius greater
than 0.1 μm contribute over 90% of the total rock permeability (Fig. 13). 6. Conclusions
The type of pore structure also has a certain relationship with the
morphology of the mercury intrusion curve. The mercury intrusion (1) The main lithologies of tight sandstones in the Shihezi Formation
curves for some samples are shown in Fig. 14a. This figure illustrates in the S block of the eastern Ordos Basin are lithic feldspar
that from the Type III pore structure sandstone to the Type I pore sandstone, lithic quartz sandstone, feldspar lithic sandstone and
structure sandstone, the mercury intrusion curve gradually deviates to lithic sandstone. The pore types include residual intergranular
the lower left (Fig. 14b). This result can be explained by a description of pores, intergranular dissolution pores, intragranular dissolution
the mercury injection and mercury removal processes. pores, intercrystalline pores and microfractures.
For the initial mercury intrusion process, there is almost no change in (2) The porosity of the rock samples ranges between 1.4% and
mercury saturation due to insufficient space at the edge of the rock to 19.5%, with an average of 9.95%; the permeability is distributed
allow mercury to enter; i.e., the mercury saturation is zero. Only when between 0.030 mD and 1.941 mD, with an average of 0.342 mD;
the mercury pressure is increased to a certain extent does a significant and the water saturation varies between 1.0% and 69.5%, with an
amount of mercury penetrate. As the mercury injection pressure in­ average of 30.7%. The water saturation of rock samples has
creases, pores with a large throat radius are filled first. These pores have negative correlations with rock porosity and permeability. This
small capillary pressures and therefore can be filled with mercury at result occurs because the low-porosity and low-permeability
lower mercury intrusion pressures. At this point, there is a significant sandstones are denser and their pore size is smaller, so more
increase in mercury saturation, and the corresponding mercury curve moisture is present in the interior of the micropores to become
deviates significantly to the lower left. Therefore, the mercury intrusion bound water or free water.
curve of the Type I pore structure sandstone deviates to the lower left (3) The Smax values of samples are distributed from 65.38 to 95.22%,
(Fig. 14b). with an average of 87.77%. Pd ranges from 0.23 to 1.46 MPa,
For the mercury removal process, mercury is first discharged from with an average of 0.61 MPa. Pm varies between 2.07 and 63.12
the smallest pores and throats with the maximum mercury intrusion MPa, with an average of 15.05 MPa rm is distributed from 0.012
pressure. Type III pore structure sandstones usually develop only pores to 0.357 μm, with an average of 0.095 μm. Both Pd and Pm have
that are connected by smaller throats, so mercury is preferentially dis­ good negative correlations with rock permeability. The median
charged. Therefore, under a small change in capillary pressure, the pore throat radius (rm) has a good positive correlation with rock
mercury saturation value of the Type III pore structure sandstone is permeability. This indicates that the larger the Pd or Pm value, the
significantly reduced, and the corresponding mercury removal curve smaller the corresponding rm value, the denser the rock and the
deviates significantly to the upper right. Correspondingly, the mercury smaller the production capacity. Large pores with a throat radius
removal curve of the Type I pore structure sandstone deviates to the

13
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

greater than 0.1 μm are most favorable for rock permeability and Han, C., Han, M., Jiang, Z.X., Han, Z.Z., Li, H., Song, Z.G., Zhong, W.J., Liu, K.X.,
Wang, C.H., 2019. Source analysis of quartz from the Upper Ordovician and Lower
usually contribute more than 90% of the total permeability.
Silurian black shale and its effects on shale gas reservoir in the Southern Sichuan
(4) Mineral and element analysis results show that the silica in the Basin and its periphery, China. Geol. J. 54 (1), 439–449.
sandstones of the target layer was not affected by hydrothermal Haneef, S., Johnson, J., Thompson, G., Wood, G., 1993. The degradation of coupled
fluid during the formation process. The sedimentary environment stones by wet deposition processes. Corrosion Sci. 34 (3), 497–510.
Harris, N.B., Miskimins, J.L., Mnich, C.A., 2011. Mechanical anisotropy in the Woodford
of the Shihezi Formation sandstone was characterized by fresh shale, Permian Basin: origin, magnitude, and scale. Lead. Edge 30, 284–291.
water, a dry-hot climate and oxidizing conditions. The major Jiang, L., Sainoki, A., Mitri, H., Ma, N., Liu, H., Hao, Z., 2016. Influence of fracture-
element combinations have good correlations with the pore induced weakening on coal mine gateroad stability. Int. J. Rock Mech. Min. Sci. 88,
307–317.
structure parameters. These results reflect that as the debris Lai, J., Wang, G., 2015. Fractal analysis of tight gas sandstones using High-Pressure
particles or quartz grains increase, the petrophysical properties Mercury Intrusion techniques. J. Nat. Gas Sci. Eng. 24, 185–196.
gradually become better. This is consistent with the gas test re­ Lai, J., Wang, G.W., Wang, S., Cao, J.T., Li, M., Pang, X.J., Zhou, Z.L., Fan, X.Q., Dai, Q.
Q., Yang, L., He, Z.B., Qin, Z.Q., 2018a. Review of diagenetic facies in tight
sults of the Shihezi Formation. That is, the gas content increased sandstones: Diagenesis, diagenetic minerals, and orediction via well logs. Earth Sci.
with increasing quartz content. Rev. 185, 234–258.
(5) According to the results of petrophysical properties, minerals, Lai, J., Wang, G.W., Wang, Z.Y., Chen, J., Pang, X.J., Wang, S.C., Zhou, Z.L., He, Z.B.,
Qin, Z.Q., Fan, X.Q., 2018b. A review on pore structure characterization in tight
elements and pore structure parameters for the study of tight sandstones. Earth Sci. Rev. 177, 436–457.
sandstones, classification criteria for pore structure are proposed. Lasaga, A., 1984. Chemical kinetics of water-rock interactions. J. Geophys. Res. 89 (B6),
4009–4025.
Li, M., Luo, J.L., Zhao, H.T., Wang, S.F., Fu, X.Y., Kang, R., 2015. Impact of the
Declaration interest competing interest
diagenetic evolution of different lithology on tight sandstone reservoir performance:
A case study from He 8 natural gas reservoir of the upper paleozoic in Eastern Ordos
We confirm that this article does not have any conflict of interest. Basin. J. NW Univ. 45 (1), 97–106 (in Chinese with English abstract).
Li, Z., Wu, S.H., Xia, D.L., Zhang, X.F., Huang, M., 2017. Diagenetic alterations and
reservoir heterogeneity within the depositional facies: a case study from
CRediT authorship contribution statement distributary-channel belt sandstone of upper triassic yanchang formation reservoirs
(ordos basin, china). Mar. Petrol. Geol. 86, 950–971.
Shuai Yin: Methodology, Writing - review & editing. Li Dong: Liu, J., Ding, W., Wang, R., Wu, Z., Gong, D., Wang, X., Yin, S., Jiao, B., 2018. Quartz
types in shale and their effect on geomechanical properties: An example from the
Investigation. Xia Yang: Investigation. Ruyue Wang: Investigation. lower Cambrian Niutitang Formation in the Cen’gong block, South China. Appl. Clay
Sci. 163, 100–107.
Acknowledgements Liu, X., Wang, J., Ge, L., Hu, F., Li, C., Li, X., Yu, J., Xu, H., Lu, S., Xue, Q., 2017a. Pore-
scale characterization of tight sandstone in Yanchang Formation Ordos basin China
using micro-CT and SEM imaging from nm- to cm-scale. Fuel 209, 254–264.
This research was supported by the Open Fund (TPR-2018-06) of Key Liu, Y., Hu, W., Cao, J., Wang, X., Tang, Q., Wu, H., 2017b. Diagenetic constraints on the
Laboratory of Tectonics and Petroleum Resources (China University of heterogeneity of tight sandstone reservoirs: a case study on the upper triassic Xujiahe
formation in the Sichuan Basin, Southwest China. Mar. Petrol. Geol. 92, 650–669.
Geosciences). Loucks, R.G., Reed, R.M., Ruppel, S.C., Jarvie, D.M., 2009. Morphology, genesis, and
distribution of nanometer-scale pores in siliceous mudstones of the Mississippian
Appendix A. Supplementary data Barnett shale. J. Sediment Res79, 848–861.
Morad, S., Al-Ramadan, K., Ketzer, J.M., 2010. The impact of diagenesis on the
heterogeneity of sandstone reservoirs: A review of the role of depositional facies and
Supplementary data to this article can be found online at https://doi. sequence stratigraphy. AAPG (Am. Assoc. Pet. Geol.) Bull. 94 (8), 1267–1309.
org/10.1016/j.jngse.2020.103189. Nabawy, B.S., G� eraud, Y., Rochette, P., Bur, N., 2009. Pore-throat characterization in
highly porous and permeable sandstones. AAPG (Am. Assoc. Pet. Geol.) Bull. 93 (6),
719–739.
References Nelson, P.H., 2009. Pore-throat sizes in sandstones, tight sandstones, and shales. AAPG
(Am. Assoc. Pet. Geol.) Bull. 93 (3), 329–340.
Abbaszadeh, M., Fujii, H., Fujimoto, F., 1996. Permeability prediction by hydraulic flow Pang, Y.M., Guo, X.W., Han, Z.Z., Zhang, X.H., Zhu, X.Q., Hou, F.H., Han, C., Song, Z.G.,
units – theory and applications. SPE Form. Eval. 11 (4), 263–271. Xiao, G.L., 2019. Mesozoic–Cenozoic denudation and thermal history in the Central
Abdullah, M., Eswaran, P., 2014. Quartz surface morphology of Tertiary rocks from Uplift of the South Yellow Sea basin and the implications for hydrocarbon systems:
North East Sarawak, Malaysis: implications for paleo-depositional environment and Constraints from the CSDP-2 borehole. Mar. Petrol. Geol. 99, 355–369.
reservoir rock quality predictions. Petrol. Explor. Dev. 41 (6), 697–704. Pittman, E.D., 1992. Relationship of porosity and permeability to various parameters
Anovitz, L.M., Cole, D.R., 2015. Characterization and analysis of porosity and pore derived from mercury injectioncapillary pressure curves for sandstone. AAPG (Am.
structures. Rev. Mineral. Geochem. 80 (1), 161–164. Assoc. Pet. Geol.) Bull. 76, 191–198.
Bloch, Salman, Lander, R.H., Bonnell, L., 2002. Anomalously high porosity and Qiu, X., Liu, C., Mao, G., Deng, Wang, F., 2014. Late Triassic tuff intervals in the Ordos
permeability in deeply buried sandstone reservoirs: origin and predictability. AAPG basin, Central China: Their epositional, petrographic, geochemical characteristics
(Am. Assoc. Pet. Geol.) Bull. 86 (2), 301–328. and regional implications. J. Asian Earth Sci. 80, 148–160.
Cao, Y.T., Liu, L., Wang, C., Kang, L., Li, D., Yang, W.Q., Zhu, X.H., 2019. Timing and Qiu, X., Liu, C., Wang, F., Deng, Y., Mao, G., 2015a. Trace and rare earth element
nature of the partial melting processes during the exhumation of the garnet–bearing geochemistry of the upper triassic mudstones in the southern Ordos Basin, Central
biotite gneiss in the southern Altyn Tagh HP/UHP belt, Western China. J. Asian China. Geol. J. 50 (4), 399–413.
Earth Sci. 170, 274–293. Qiu, X., Liu, C., Mao, G., Deng, Y., Wang, F., Wang, J., 2015b. Major, trace and platinum-
Chang, X.C., Wang, Y., Shi, B.B., Xu, Y.D., 2019. Charging of Carboniferous volcanic group element geochemistry of the upper triassic nonmarine hot shales in the Ordos
reservoirs in eastern Chepaizi Uplift, Junggar Basin (NW China) constrained by oil Basin, central china. Appl. Geochem. 53, 42–52.
geochemistry and fluid inclusion. AAPG (Am. Assoc. Pet. Geol.) Bull. 103, 1–14. Rezaee, R., Saeedi, A., Clennell, B., 2012. Tight gas sands permeability estimation from
https://doi.org/10.1306/12171818041. mercury injection capillary pressure and nuclear magnetic resonance data. J. Petrol.
Dou, W.T., Liu, X.S., Wang, T., 2010. The origin of formation water and the regularity of Sci. Eng. 88 89, 92–99.
gas and water distribution for the Sulige gas field, Ordos basin. Acta Pet. Sin. 31 (5), Sakhaee-Pour, A., Bryant, S.L., 2014. Effect of pore structure on the producibility of
767–772 (in Chinese with English abstract). tight-gas sandstones. AAPG Bull. 98, 663e694.
Du, Q.X., Han, Z.Z., Shen, X.L., Han, C., Song, Z.G., Gao, L.H., Han, M., Zhong, W.J., Shanley, K.W., Cluff, R.M., 2015. The evolution of pore-scale fluid-saturation in low
2019. Geochronology and geochemistry of Permo-Triassic sandstones in eastern Jilin permeability sandstone reservoirs. AAPG Bull. 99, 1957–1990.
Province (NE China): implications for final closure of the Paleo-Asian Ocean. Geosci. Shi, B.B., Chang, X.C., Yin, W., Li, Y., Mao, L.X., 2019. Quantitative evaluation model of
Front. 10, 685–706. https://doi.org/10.1016/j.gsf.2018.03.014. tight sandstone reservoirs based on statistical methods - A case study of the Triassic
Dutton, S.P., Loucks, R.D., 2010. Diagenetic controls on evolution of porosity and Chang 8 tight sandstones, Zhenjing area, Ordos Basin, China. J. Petrol. Sci. Eng. 173,
permeability in lower Tertiary Wilcox sandstones from shallow to ultradeep 601–616.
(200–6700m) burial, Gulf of Mexico Basin, U.S.A. Mar. Petrol. Geol. 27, 69–81. Torabi, A., Fossen, H., Braathen, A., 2013. Insight into petrophysical properties of
Fu, N., Yang, S.C., He, Q., Xu, H., Lin, Q., 2016. High-efficiency reservoir formation deformed sandstone reservoirs. AAPG Bull. 97 (4), 619–637.
conditions of tigh sandstone gas in Linxing-Shenfu blocks on the east margin of Wang, G., Chang, X., Yin, W., Li, Y., Song, T., 2017. Impact of diagenesis on reservoir
Ordos Basin. Acta Pet. Sin. 37 (Suppl. 1), 111–119 (in Chinese with English abstract). quality and heterogeneity of the upper triassic chang 8 tight oil sandstones in the
Ghanizadeh, G., Clarkson, C.R., Aquino, S., Ardakani, O.H., Sane, H., 2015. Petrophysical Zhenjing area, Ordos Basin, China. Mar. Petrol. Geol. 83, 84–96.
and geomechanical characteristics of Canadian tight oil and liquid-rich gas Wang, S.J., Li, X.P., Schertl, H.P., Feng, Q.D., 2019. Petrogenesis of early cretaceous
reservoirs: I. pore network and permeability characterization. Fuel 153, 664–681. andesite dykes in the Sulu orogenic belt, eastern China. Mineral. Petrol. 113 (1),
77–97.

14
S. Yin et al. Journal of Natural Gas Science and Engineering 76 (2020) 103189

Xi, K., Cao, Y., Haile, B.G., Zhu, R., Jahren, J., Bjørlykke, K., Zhang, X., Hellevang, H., Yin, S., Gao, Z., 2019. Numerical study on the prediction of “sweet spots” in a low
2016. How does the pore-throat size control the reservoir quality and oiliness of tight efficiency-tight gas sandstone reservoir based on a 3D strain energy model. IEEE
sandstones? The case of the Lower Cretaceous Quantou Formation in the southern Access 7, 1–12.
Songliao Basin, China. Mar. Petrol. Geol. 76, 1–15. Yin, S., Zhao, J., Wu, Z., Ding, W., 2018a. Strain energy density distribution of a tight gas
Xiao, D., Lu, Z., Jiang, S., Lu, S., 2016. Comparison and integration of experimental sandstone reservoir in a low-amplitude tectonic zone and its effect on gas well
methods to characterize the full-range pore features of tight gas sandstone—a case productivity: A 3D FEM study. J. Petrol. Sci. Eng. 170 (6), 1–16.
study in Songliao basin of China. J. Nat. Gas Sci. Eng. 34, 1412–1421. Yin, S., Lv, D., Ding, W., 2018b. New method for assessing microfracture stress sensitivity
Yang, R., Jin, Z., Van Loon, A., Han, Z., Fan, A., 2017. Climatic and tectonic controls of in tight sandstone reservoirs based on acoustic experiments. Int. J. GeoMech. 18 (4),
lacustrine hyperpycnite origination in the late triassic Ordos Basin, Central China: 1–16.
implications for unconventional petroleum development. AAPG (Am. Assoc. Pet. Yin, S., Lv, D., Jin, L., Ding, W., 2018c. Experimental analysis and application of the
Geol.) Bull. 101, 95–117, 01. effect of stress on continental shale reservoir brittleness. J. Geophys. Eng. 15 (2),
Yin, S., Tian, T., Wu, Z., 2019a. Developmental characteristics and distribution law of 478–494.
fractures in a tight sandstone reservoir in a low-amplitude tectonic zone, eastern Yin, S., Jia, Q., Ding, W., 2018d. 3D paleotectonic stress field simulations and fracture
Ordos Basin, China. Geol. J. 54 (6), 1–16. prediction for marine-continental transitional facies forming a tight-sandstone
Yin, S., Han, C., Wu, Z., Li, Q., 2019b. Developmental characteristics, influencing factors reservoir in a highly deformed area. J. Geophys. Eng. 15 (4), 1214–1230.
and prediction of fractures for a tight gas sandstone in a gentle structural area of the Zeng, P., Zhao, J., Li, Z., Guo, J., Yang, M., 2005. Experimental study concerning the
Ordos Basin, China. J. Nat. Gas Sci. Eng. 72, 1–14. effect of temperature, effective stress and water saturation on the permeability of
Yin, S., Xie, R., Wu, Z., Liu, J., Ding, W., 2019c. In situ stress heterogeneity in a highly tight sandstone. Nat. Gas Geosci. 16 (1), 31–34 (in Chinese with English abstract).
developed strike-slip fault zone and its effect on the distribution of tight gases: A 3D Zhang, X., Feng, Q., Sun, P., Li, W., 2010. Characteristics of high gamma ray reservoir of
finite element simulation study. Mar. Petrol. Geol. 81 (1), 1–17. Yanchang formation in Ordos Basin. Chin. J. Geophys. 53 (1), 205–213.
Yin, S., Ding, W., 2019. Evaluation indexes of coalbed methane accumulation in the Zou, C., Zhu, R., Liu, K., Su, L., Bai, B., Zhang, X., Yuan, X., Wang, J., 2012. Tight gas
strong deformed strike-slip fault zone considering tectonics and fractures: A 3D sandstone reservoirs in China: characteristics and recognition criteria. J. Petrol. Sci.
geomechanical simulation study. Geol. Mag. 156 (6), 1052–1068. Eng. 88–89, 82–91.

15

You might also like