Evaluation of nanoscale accessible pore structures for improved

Nov 28, 2018 - Evaluation of nanoscale accessible pore structures for improved prediction of gas production potential in Chinese marine shales. Yang W...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of Winnipeg Library

Fossil Fuels

Evaluation of nanoscale accessible pore structures for improved prediction of gas production potential in Chinese marine shales Yang Wang, Yong Qin, Rui Zhang, Lilin He, Lawrence M. Anovitz, Markus Bleuel, David F.R. Mildner, Shimin Liu, and Yanming Zhu Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.8b03437 • Publication Date (Web): 28 Nov 2018 Downloaded from http://pubs.acs.org on December 1, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1 2

Evaluation of nanoscale accessible pore structures for improved prediction of gas production potential in Chinese marine shales

3 4

Yang Wanga,b, Yong Qina,b*, Rui Zhangc,d **, Lilin Hed, Lawrence M. Anovitze, Markus Bleuelf,g, David F. R. Mildnerf, Shimin Liuc, Yanming Zhua,b aKey

5 6 7 8 9 10 11 12 13 14 15 16

Laboratory of Coalbed Methane Resources and Reservoir Formation Process, China University of Mining and Technology, Xuzhou, Jiangsu 221008, China bSchool of Resources and Earth Science, China University of Mining and Technology, Xuzhou, Jiangsu 221116, China cDepartment of Energy and Mineral Engineering, G3 Center and Energy Institute, The Pennsylvania State University, University Park, PA 16802, USA dNeutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, TN 37830, USA eChemical Sciences Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA fNIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA gDepartment of Materials Science and Engineering, A. James Clark School of Engineering, University of Maryland, College Park, MD 20742, USA

17 18

Corresponding authors: *Email: [email protected] ; **Email: [email protected]

19

Abstract

20

The Lower Cambrian Niutitang and Lower Silurian Longmaxi shales in the Upper Yangtze

21

Platform (UYP) are the most promising strata for shale gas exploration in China. Knowledge of

22

the nanoscale pore structure may improve the prediction of the gas production potential in Chinese

23

marine shales. A systematic investigation of the pore accessibility and its impact on methane

24

adsorption capacity has been conducted on shale samples using various techniques including

25

geochemical and mineralogical analyses, field-emission scanning electron microscopy (FE-SEM),

26

small-angle neutron scattering (SANS), helium porosimetry and methane adsorption. The results

27

show that organic matter (OM) pores with various shapes dominate the pore systems of these

28

shales. OM tended to mix with clay minerals and converted to organo-clay complexes, developing

29

plentiful micro- and mesopores. A unified fit model with two pore structures, fractal pores and

30

finite pores, was used to model the SANS data to characterize the pore structure of the shales. Both

31

mass and surface fractals are identified for each pore structure. Total porosity estimated by the 1 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 44

32

Porod method ranges between 2.35 % and 16.40 %, of which the porosity for finite pores ranges

33

between 0.35 % and 6.36 %, and the porosity for the fractal pores ranges between 2.07 % and 8.51

34

%. The fraction of open pores was evaluated by comparing the porosities estimated by He

35

porosimetry and SANS. We find that the fraction of open pores is higher than 64 % for most of

36

these shales. Correlation analyses suggest that clay and total organic carbon (TOC) have opposite

37

effects on pore structure and methane adsorption capacity. Samples with higher clay contents have

38

higher pore accessibility and lower total porosity, surface area and maximum methane adsorption,

39

whereas samples with higher TOC content show the inverse relationships. The high percentage of

40

open pores may reduce methane adsorption capacity in these shales, whereas low pore accessibility

41

may reduce methane production at specific pressure differences. Thus, both TOC and pore

42

accessibility may be essential controlling factors in methane production from shale gas reservoirs.

43

Keywords

44

Pore structure; pore accessibility; methane adsorption; Chinese marine shale; Upper Yangtze

45

Platform

46

1. Introduction

47

In response to growing worldwide energy demands, shale gas is becoming an increasingly

48

valuable energy resource. Following the shale gas revolution in North America and the

49

development of the techniques of horizontal drilling and hydraulic fracturing, intensive efforts in

50

shale gas explorations have been conducted in China [1-3]. Unlike conventional gas reservoirs,

51

“shale gas reservoirs” are self-contained source reservoir systems that are generally characterized

52

by high total organic carbon (TOC) contents compared to conventional gas reservoirs, complex

53

mineral compositions, low porosities and low permeability matrices with extensive, heterogeneous,

54

nanoscale pores [4-7]. This nanoscale pore structure is one of the key factors influencing the 2 ACS Paragon Plus Environment

Page 3 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

55

occurrence, storage capacity, and flow behavior of gas in these reservoirs [8-15]. Shale gases occur

56

as both adsorbed and free gas and are thus directly affected by the nanoscale pore structures.

57

Quantitative characterization of this nanoscale pore structure is, therefore, essential for evaluating

58

gas production potential and providing clues to increasing gas extraction efficiency [9, 11, 16, 17].

59

Due to the diversity, complexity and heterogeneity of shale pore structures, it is difficult to

60

characterize pore structures using a single technique [11, 18-20]. A variety of techniques have,

61

therefore, been used to characterize the pore structures of shales, in particular their porosity, pore

62

volume and specific surface area. Direct imaging methods, including field emission-scanning

63

electron microscopy (FE-SEM), transmission electron microscopy (TEM) and focused ion beam-

64

scanning electron microscopy (FIB-SEM), have been successfully employed to observe the pore

65

morphology and connectivity of pore networks [21-24]. However, these direct imaging techniques

66

assess only limited areas of a sample at a given time, and thus their statistical reliability is limited.

67

Indirect methods, including helium porosimetry, mercury intrusion porosimetry and low-pressure

68

N2/CO2 adsorption, are also commonly used to quantitatively characterize the pore size distribution,

69

pore volume and specific surface area of shales; however, these techniques typically interrogate

70

larger sample volumes and can only characterize accessible pores and those with pore diameters

71

larger than that of the probe molecule. Thus, adsorption techniques are unable to provide

72

information about closed porosity [4, 19, 25-27]. In contrast, small-angle neutron scattering

73

(SANS), a nondestructive physical technique, can be used to characterize both accessible and

74

closed pores since it records the scattering intensity from all pores within the detectable size range

75

[28-38]. Closed pores may play an essential role in gas shale production because they may contain

76

methane, and thus the efficiency of gas recovery may depend on the extent to which they can be

77

accessed. Given the apparent strengths and weaknesses of each of these approaches, therefore, a

3 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 44

78

combination of different characterization techniques should lead to a more precise prediction of

79

gas production potential.

80

The marine shales in the Upper Yangtze Platform (UYP), the Lower Silurian Longmaxi

81

Formation and the Lower Cambrian Niutitang Formation, are the primary targets for shale gas

82

exploration and development in China because of their widespread occurrence, large thicknesses,

83

organic richness, and favorable mineral compositions [1, 3, 13, 39]. Compared to the Longmaxi

84

shales, the Niutitang shales has been more deeply buried and has higher thermal maturity [3,7]. In

85

addition, high initial gas production rates of (10-50)×104 m3 per day have been reported for most

86

pilot wells in the Jiaoshiba area of the Longmaxi shales [40, 41]. The breakthrough discovery of

87

this shale gas field represents a significant milestone for shale gas exploration in China [3, 41].

88

In order to provide a more thorough assessment of the pore structures and gas production

89

potential of the UYP, a series of pore structure characterization techniques were applied to

90

characterize pore morphologies, porosities, pore volumes and specific surface areas of the marine

91

shales from the Lower Cambrian Niutitang and Lower Silurian Longmaxi Formations. In this study,

92

we have used a combination of FE-SEM, SANS, helium porosimetry and high-pressure methane

93

adsorption to investigate pore accessibility and its effect on methane adsorption. Uniquely, the

94

correlations between pore accessibility and methane adsorption properties indicating that pore

95

accessibility could be an indicator for methane production from shale gas reservoirs. The results

96

will be broadly applicable to shale gas exploration and development including reservoir

97

assessment and evaluation, gas-in-place estimation, carbon sequestration and other characteristics

98

of the UYP.

4 ACS Paragon Plus Environment

Page 5 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

99 100

Energy & Fuels

2. Materials and methods 2.1 Sample collection

101

The UYP is located in the western portion of the Yangtze Platform in southern China (Fig. 1).

102

It includes the Chongqing city, northern Guizhou and Yunnan, western Hubei and Hunan, and

103

eastern Sichuan province. Geological studies indicate that the Niutitang Formation and the

104

Longmaxi Formation formed on a neritic shelf with a stagnant marine sedimentary environment

105

[1, 3]. These marine shales are currently the most prominent exploration and production targets

106

for shale gas in China.

107 108

Fig. 1 Geographical map of UYP and sampling location

109

A total of six shale samples were analyzed in this study. Four fresh Longmaxi Formation shale

110

core samples were collected from the WX2 well, located in northeast Chongqing near the edge of

111

the Sichuan Basin. Two fresh Niutitang Formation shale core samples were collected from the 5 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 44

112

QD1 well located in northeast Yunnan near the southwestern edge of the UYP. Both of these shales

113

are widely distributed around the UYP. Although both are thermally over-mature, the thermal

114

maturity of the Niutitang shales is much greater than that of the Longmaxi shales, which indicates

115

that the Niutitang experienced more severe compaction and underwent more complicated

116

diagenesis than the Longmaxi [1].

117

2.2 Geochemical and mineralogical tests

118

Geochemical parameters and mineral compositions were characterized for all the shale samples

119

measured in this study. Analysis of TOC content was performed using a LECO CS-230 carbon

120

and sulfur analyzer. Due to the lack of vitrinite in the Lower Paleozoic shales, the reflectance of

121

pyrobitumen was measured using a Leica MPV-SP microphotometer. The average pyrobitumen

122

reflectance was then converted to an equivalent vitrinite reflectance (EqVRr) using the equation of

123

Schoenherr et al. (2007), i.e., EqVRr =(BRr+0.2443)/1.0495, where BRr is the pyrobitumen

124

reflectance [42]. The mineralogy of each sample was obtained using a RIGAKU D/Max-3B type

125

X-ray diffractometer (XRD). Preparation, analysis and interpretation procedures were conducted

126

following the Chinese Oil and Gas Industry Standard SY/T 5163-2010. Multicomponent

127

quantifications were achieved by Rietveld refinement using the TOPAS software package. All

128

analyses were carried out at the Experimental Research Center of East China Branch, SINOPEC.

129

2.3 FE-SEM

130

FE-SEM observations were performed using a Quanta 200F instrument at the China University

131

of Petroleum, Beijing Campus (CUP). The FE-SEM images provide a spatial resolution of 1.2 nm,

132

which is sufficient to characterize the nanoscale pore morphology. Before the experiment, the

133

samples were pre-polished using an argon ion mill to create an artifact-free surface. Au was then

134

coated onto the artifact-free surface to improve the image quality by enhancing the electrical 6 ACS Paragon Plus Environment

Page 7 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

135

conductivity. The experiment and subsequent analysis were carried out at a constant temperature

136

of 24 °C and a constant humidity of 35 %. High-resolution images were obtained at an accelerating

137

voltage of 10 kV to 15 kV with a working distance of 4 mm to 5 mm.

138

2.4 SANS

139

Of the six samples analyzed by SANS in this study, the first two (WX2-8 and WX2-33) were

140

analyzed as powders and the other four (QD1-L3, QD1-L4, WX2-49 and WX2-54) as disks.

141

Analysis of the powdered shales was conducted using the general-purpose small-angle neutron

142

scattering diffractometer (GP-SANS) at the High Flux Isotope Reactor (HFIR) at Oak Ridge

143

National Laboratory (ORNL) [43]. The samples had a non-uniform particle size of < 0.5 mm, and

144

an overall sample thickness of 1.6 mm. These samples were measured under an ambient condition

145

with sample-to-detector distances of 1.1 m, 8.8 m and 19.2 m. The neutron wavelength 𝜆 was set

146

at 4.75 Å for the 1.1 m and 8.8 m measurements, and 12 Å for the 19.2 m measurements to extend

147

the scattering vector 𝑄 range to values as small as possible (the largest sizes possible), where 𝑄 = 4

148

𝜋sin 𝜃/𝜆 and 𝜃 is half of the neutron scattering angle. The overall 𝑄 range was between ~1.5×10-3

149

Å-1 and ~0.8 Å-1, which corresponds to pore sizes between ~3 Å and ~1.6×103 Å using the

150

empirical relation 𝑅 ≈ 2.5/𝑄 [44]. The absolute scattering intensities were obtained using

151

measured sample transmission and thickness values as well as a standard scattering intensity

152

measured on porous silica [45]. All the 1D scattering curves were obtained using the NCNR Igor

153

macros package [46].

154

In order to reduce multiple scattering, especially at low 𝑄 and long wavelengths [28, 47], the

155

remaining samples were analyzed as thin disks with an arbitrary orientation, using the method

156

outlined by Anovitz et al. [28]. These analyses were performed using the NG-7 30m small-angle

157

neutron scattering diffractometer (NG7-SANS) at the NIST Center for Neutron Research (NCNR) 7 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 44

158

at the National Institute of Standards and Technology (NIST) [48]. Four thin disk shale samples

159

were commercially polished to a thickness of 0.1 mm and mounted on 1 mm thick quartz glass

160

slides [28, 47]. These samples were analyzed with the normal to the incident beam and under

161

ambient condition. Four SANS geometries were used: 1 m and 4 m sample-to-detector distances,

162

and a 13 m distance with and without a set of biconcave MgF2 lenses. The neutron wavelength

163

was set at 6 Å for SANS geometries of 1 m, 4 m and 13 m, and at 8.1 Å for the configuration with

164

the lenses. These settings cover scattering vector 𝑄 range between ~1×10-3 Å-1 and ~0.6 Å-1, which

165

corresponds to a pore size range between ~4 Å and ~2.5×103 Å.

166

2.5 Porosity using helium porosimetry

167

The total accessible porosity of the shale samples was determined at the Experimental Research

168

Center of East China Branch, SINOPEC using a technique based on Boyle's Law, by measuring

169

the grain volume under ambient conditions and bulk volume by Archimedes’ principle with helium

170

porosimetry [47]. Samples were oven dried at 115 °C to a constant mass (±1 mg). Porosity was

171

calculated from the difference between bulk volume and grain volume.

172

2.6 High-pressure methane adsorption

173

High-pressure methane adsorption isotherms were measured at a constant temperature of 40 °C

174

and pressure up to 9 MPa using a volumetric sorption apparatus (IS-300 isothermal

175

adsorption/desorption analyzer) manufactured by Terratek, USA. Before the adsorption

176

measurement, samples were crushed to a particle size of less than 250 µm (60 mesh) and dried for

177

24 h at 105 °C. The detailed methane adsorption measurement procedure is described by Ji et al.

178

[49]. Methane sorption data were normalized to sample weight and parametrized using the

179

Langmuir monolayer adsorption model [50].

8 ACS Paragon Plus Environment

Page 9 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

180

3. Results and discussion

181

3.1 Geochemical and mineralogical parameters

182

Detailed geochemical and mineralogical characteristics of the marine shale samples are listed

183

in Table 1. Mass fraction is used for the chemical compositions. TOC contents range between 1.52

184

wt.% and 6.33 wt.%. The marine shales in UYP have experienced strong thermal maturation and

185

reached the post-mature stage with EqVRr values ranging from 2.06 % to 3.03 %.

186

XRD analysis demonstrates that both Cambrian Niutitang shales and Silurian Longmaxi shales

187

have complex mineralogy (Table 1 and Fig. 2). Quartz and clay are the major components with an

188

average of 49.3 wt.% (38.0–56.6 wt.%) and 28.9 wt.% (22.0–41.0 wt.%), respectively. Fig. 2a

189

shows that these shales also contain pyrite, calcite, dolomite, K-feldspar and plagioclase. Illite is

190

the most abundant clay mineral ranging between 10.6 wt.% and 26.3 wt.% with a mean value of

191

17.4 wt.%. Fig. 2b shows the other clay minerals including smectite, chlorite and illite/smectite

192

mixed layer clays.

9 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 44

193 194

Fig. 2 Chemical composition of the six Chinese marine shales: (a) Total mineral contents; (b) Clay mineral contents

10 ACS Paragon Plus Environment

Page 11 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Energy & Fuels

195

Table 1 Chemical composition of the six Chinese marine shales Formation

Niutitang (Є1)

Longmaxi (S1)

Sample ID

Quartz (wt.%)

K-feldspar (wt.%)

Plagioclase (wt.%)

Dolomite (wt.%)

Calcite (wt.%)

Pyrite (wt.%)

Clay (wt.%)

Illite (wt.%)

Smectite (wt.%)

Chlorite (wt.%)

Illite/Smectite mixed layer (wt.%)

TOC (wt.%)

EqVRr (wt.%)

QD1-L3

46.61

4.64

7.25

1.35

5.61

1.64

29.59

15.09

5.03

1.55

7.93

3.29

2.85

QD1-L4

54.15

1.44

3.64

/

6.98

2.39

27.08

19.61

1.15

0.77

5.55

4.32

3.03

WX2-8

37.42

1.48

14.28

/

2.17

2.76

40.38

25.90

/

5.12

9.36

1.52

2.06

WX2-33

46.33

/

9.60

2.94

13.71

3.82

21.54

10.38

/

2.55

8.62

2.05

2.12

WX2-49

49.41

1.15

5.27

/

3.26

9.10

27.58

16.09

1.53

3.35

6.61

4.25

2.26

WX2-54

50.77

3.09

8.43

1.12

3.28

5.99

20.98

13.58

1.12

2.15

4.12

6.33

2.32

196

11 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

197

Page 12 of 44

3.2 Qualitative pore morphology characteristics using FE-SEM

198

Pore structure has an enormous impact on the occurrence of shale gas, the calculation of gas

199

in place (GIP) and hydraulic fracturing. Qualitative analysis of pore morphology helps reveal the

200

origin of nanoscale pores and the nature of microscopic pore structures. Based on detailed pore

201

classifications proposed in previous studies [6, 9, 51-52], we have divided the pores observed in

202

our marine shales into four categories: Interparticle (interP) pores, intraparticle (intraP) pores,

203

organic matter (OM) pores and microfractures.

204

FE-SEM observations (Fig. 3) suggest that OM pores dominate the pore system of the marine

205

shales. Their diameters vary from several to hundreds of nanometers and they display various

206

shapes such as spherical to ellipsoidal, schistose-like, irregular polygons and honeycombs (Figs.

207

3a, b and c). OM pores are well-developed and often form pore networks with functional

208

connectivity. Since these shales have reached the over-mature stage, it is likely that OM pores

209

were formed gradually during thermal evolution. Furthermore, OM does not, necessarily, exist in

210

isolation. OM is commonly intergrown with pyrite framboids (Figs. 3b and c), forming a large

211

number of nanoscale pores (pore size in the nanometer range). OM also tends to partially mix with

212

clay minerals (Fig. 3a), becoming converted to organo-clay complexes with a full range of micro-

213

/mesopores sizes (pore size < 50 nm).

214

IntraP pores are another major contributor to the porosity of these marine shales. This type of

215

pore is mainly associated with pyrite framboids, clay aggregates and dissolution of unstable

216

minerals such as plagioclase and calcite (Fig. 3f). Pyrite framboids were observed in these shales,

217

and nanoscale pores were commonly observed between individual pyrite crystals. These were

218

probably formed by incomplete cementation or dissolution of the pyrite grains (Figs. 3d and g).

219

Fig. 3e shows an example of intraP pores within clay mineral aggregates displaying a “card house”

12 ACS Paragon Plus Environment

Page 13 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

220

structure [53]. This provides interconnected pores which may serve as essential pathways

221

enhancing gas permeability.

222 223

Fig. 3 FE-SEM images of the six Chinese marine shales: (a) OM pores and micro-/mesopores developed within

224

organo-clay complexes (sample WX2-54); (b) OM is commonly intergrown with pyrite framboids (sample WX2-33);

225

(c) Nanoscale pores within OM filled between pyrite framboids (sample WX2-33); (d) IntraP pores and interP pores

226

developed between the contact areas of pyrite and clay minerals (sample WX2-49); (e) IntraP pores developed within

227

clay mineral aggregates (sample QD1-L3); (f) Dissolution pores and fractures developed within calcite (sample WX2-

228

8); (g) IntraP pores developed within pyrite framboids (sample QD1-L4); (h) Microfractures developed within clay

229

minerals and OM (sample WX2-49); (i) Microfractures are commonly connected nanoscale pores (sample QD1-L4)

13 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 44

230

In contrast to OM and intraP pores, interP pores are susceptible to mechanical compaction and

231

difficult to preserve. InterP pores are usually developed at the points of contact between particles.

232

They exist between the contact areas of OM and clay minerals and along the pyrite and quartz

233

grains (Figs. 3d and i). The distribution of interP pores observed in these samples was quite

234

heterogeneous with low connectivity.

235

Various

kinds of microfractures were observed in the tested shales. Some microfractures were

236

observed within unstable minerals and appeared to be formed as a result of dissolution (Fig. 3f).

237

Microfractures within OM are usually related to hydrocarbon generation (Fig. 3h). Some

238

microfractures were also developed within clay minerals (Fig. 3h). Typically, these are 20 nm to

239

100 nm wide and about 3 μm long. Note that the origin of the microfractures in clay minerals

240

remains uncertain which is likely formed due to clay shrinkage, either during diagenesis or perhaps

241

secondarily as a result of core sample recovery and processing. Meanwhile, nanoscale pores and

242

microfractures are also influenced by mechanical compaction since both Longmaxi and Niutitang

243

Formations experienced multi-phase tectonic movements. Fig. 3i shows that microfractures

244

connect micropores (pore size < 2 nm) and macropores (pore size > 50 nm), which can improve

245

the connectivity of the pore networks. Therefore, microfractures not only provide space for free

246

and adsorbed gas to accumulate but may serve an important role in shale gas migration.

14 ACS Paragon Plus Environment

Page 15 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

247

3.3 Quantitative pore structure characteristics using SANS

248

3.3.1 Experimental scattering intensities

249 250

Fig. 4 Comparison of background-subtracted experimental scattering intensities for the six Chinese marine shales: (a)

251

log (𝐼(𝑄)) vs. log (𝑄); (b) log (𝑄4𝐼(𝑄)) vs. log (𝑄)

252

Fig. 4 shows the background-subtracted 1D scattering profiles on logarithmic scales for the six

253

Chinese marine shale samples tested. In each case, the log of the intensity decreases nearly linearly

254

with increasing log (𝑄) over the entire measured 𝑄 region, although careful examination by

255

plotting log (𝑄4𝐼(𝑄)) as a function of log (𝑄) shows that this is not completely true, especially for

256

samples WX2-8 and WX2-33. This scattering profile is typical for shales in this 𝑄 range [10, 11,

257

34, 35, 37, 38, 54-60]. The scattering intensities measured for the two Niutitang samples and one

258

of the Longmaxi samples, QD1-L3, QD1-L4 and WX2-49 (all disk samples), were similar over the

259

measured 𝑄 range, indicating that the total pore volumes and pore volume distributions in this size

260

range (between ~4 Å and ~2.5×103 Å) among these samples are quite similar. The scattering

261

intensity of WX2-54 is the highest among the tested samples over the entire 𝑄 range and slope of

262

the scattering curve is shallower. This indicates that the total pore volume of WX2-54 is greater

263

than the other five samples over this specific size range, and that this difference is probably the

264

largest at smaller pore sizes. At the opposite extreme, the scattering intensity of WX2-8 is smaller 15 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 44

265

than that of other five samples at 𝑄 > ~0.01 Å-1, implying that this sample has a smaller total pore

266

volume over this pore range compared to the others. A hump is present at low 𝑄 for samples WX2-8

267

and WX2-33, and there is a broad hump in the data for WX2-54 around 0.03 Å-1 (~8 nm). A small

268

hump in the scattering profile was also observed for samples QD1-L4, WX2-8 and WX2-49 around

269

0.2 Å-1. Similar results have been seen in both SANS and SAXS data for shales [11], coals [10,

270

61, 62], and sandstones [29, 30]. Interestingly, a hump that is typically observed in the SAXS

271

profiles of coal samples is absent in the SANS profiles for the same samples [61]. Thus, these may

272

be caused by the interlayer spacing of clays in QD1-L4, WX2-8, WX2-49 and WX2-54.

273

3.3.2 Pore structure determination

274

In order to quantitatively characterize the nanopore structure of the tested shales we have

275

applied an empirical SANS model that combines Guinier and power-law equations to fit the

276

experimental SANS data. In this unified model the scattering intensity 𝐼(𝑄) is expressed as [63]:

{

𝑛 ∑𝑖 = 1

𝑄𝑅g



𝑄2𝑅g2 𝑖 3

[

( ( )) erf

6

𝑖

3

𝛼𝑖

]}

277

𝐼(𝑄) =

278

where 𝑄 is the scattering vector; 𝐺 is the Guinier coefficient, which is equal to the scattering

279

intensity at 𝑄 = 0; 𝑅g is a radius of gyration representing a domain size; 𝐵 is the power law

280

coefficient; 𝛼 is the power law exponent; 𝑖 is the number of the pore structure, as several

281

independent pore structures can be fitted with this model; 𝑛 is the total number of pore structures

282

considered; and 𝐵𝑘𝑔 is the scattering background. Based on the experimental scattering data

283

shown in Fig. 4, the presence of power-law scattering at low and medium 𝑄 implies the existence

284

of a fractal pore structure in the tested shales, with the exceptions of WX2-8 and WX2-33.

285

Furthermore, both the deviation in the scattering intensity from power-law scattering at high 𝑄 and

𝐺𝑖𝑒

+ 𝐵𝑖

𝑄

+𝐵𝑘𝑔

(1)

16 ACS Paragon Plus Environment

Page 17 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

286

the “humps” may reflect additional fractal pore structures with finite size ranges. Thus, a model

287

with two pore structures based on Eq. 1 (with 𝑛 = 2) has been used to fit the data for the shale

288

samples. The first reflects the power-law scattering for a fractal pore structure (fractal pores); the

289

second is fitted with a combined Guinier and power-law model, and reflects a fractal pore structure

290

with a finite pore range (finite pores).

291

Fitting Eq. 1 to the scattering data for each shale sample was performed using the Irena

292

software package [64]. The results are shown in the supplementary Fig. S1 and detailed model

293

parameters are listed in the Supplementary Tables. Note that, because the thickness of the samples

294

led to excess multiple scattering at low 𝑄 for samples WX2-8 and WX2-33, the low 𝑄 cutoff of

295

model fitting was chosen at the linear range in the log (𝐼(𝑄)) ― log (𝑄) plot. The modeled

296

scattering intensity was then extrapolated to the experimental low 𝑄 limit for each sample as shown

297

in Figs. S1c and d. When the value of the fitted power law exponent 𝛼 is between 3 and 4 the data

298

may be interpreted as a surface fractal with dimensionality 𝐷s, given by 𝐷s = 6 ― 𝛼, so that 𝐷s lies

299

between 2 and 3. When the value of 𝛼 is smaller than 3 the correlation 𝐷m = 𝛼 was used to estimate

300

the mass fractal dimension 𝐷m [47, 65].

301

The estimated fractal dimensions for these shale samples are shown in Table 2. For the finite

302

pores the model yields mass fractal dimensions between 1.94 and 2.68 except for sample QD1-L4,

303

for which suggests a surface fractal dimension of 2.85. For the fractal pores, samples QD1-L4,

304

WX2-49 and WX2-54 have mass fractal dimensions 𝐷m between 2.84 and 2.93, whereas samples

305

QD1-L3, WX2-8 and WX2-33 generate surface fractal dimensions 𝐷s ranging between 2.42 and

306

2.78. The sample preparation procedure could be the reason for the existence of both mass and

307

surface fractals in the shales tested. The unified model suggests that the pore structure of the finite

308

pores has a maximum pore size ranging between ~3.8 nm and ~7.8 nm assuming spherical pore 17 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 44

309

shape, where 𝑅g = 3/5𝑅, as the radius of gyration 𝑅g ranges between ~3.0 nm and ~6.1 nm

310

(Tables S2-S7 in the supplementary). This represents the maximum domain size of this structure,

311

which is within the meso-/micropore size range. The results also indicate that the shale pores have

312

different fractal characteristics at different pore sizes.

313

Both the total porosity and the porosities of each pore structure were estimated using the Porod

314

invariant method. The Porod invariant 𝑄INV can be expressed as [66]:

315

𝑄INV = ∫𝑄max𝑄2𝐼(𝑄)𝑑𝑄 = 2𝜋2(𝜌s∗ ― 𝜌p∗ ) ∅(1 ― ∅)

316

where 𝜌s∗ and 𝜌p∗ are the scattering length densities (SLDs) of the solid matrix and pores,

317

respectively; 𝑄max and 𝑄min are the experimental maximum and minimum 𝑄 values, respectively;

318

and ∅ is the porosity. Here we assume that the SLD is the same for the fractal and finite pores for

319

porosity estimation. Nonetheless, if one were primarily OM filled, or one contrasted with a

320

different mineral than the other, the SLDs could be different, although probably not very different.

321

The value of ∅ estimated for each pore structure and the total porosity for all samples are shown

322

in Table 2. Note that the 𝑄 range for porosity estimation was confined between ~1.5×10-3 Å-1 (for

323

𝑄min) and ~0.5 Å-1 (for 𝑄max) for the reason noted above. Thus, the pore size range for total pores

324

is between ~0.5 nm and ~1.6×102 nm. For the divided two pore structures, fractal pores have the

325

same pore size range as the total pores. However, the pore size range of finite pores is sample

326

dependent. The lowest pore size limit is ~0.5 nm, while the highest pore size limit is between ~3.8

327

nm and ~7.8 nm for all the samples. The estimated total porosity ranged from 2.35 % to 16.40 %.

328

The porosity of finite pores ranged between 0.35 % and 6.36 %, and that of fractal pores between

329

2.07 % and 8.51 %. The reason why the total porosity is not the exact addition from the porosities

330

of fractal and finite pores is that the experimental scattering intensity is used to estimate ∅ total

331

while the modeled scattering intensity is used to estimate ∅ for each pore structure. Another reason

𝑄

2

min

(2)

18 ACS Paragon Plus Environment

Page 19 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

332

could be the assumption of equal SLDs for the total, fractal and finite pores, which may be different.

333

With the exception of sample QD1-L3 the pore volume of the finite pores is generally smaller than

334

that of the fractal pores.

335

Table 2 Fractal dimension 𝐷 and porosity ∅ for the six Chinese marine shales

Formation

Niutitang (Є1)

Longmaxi (S1)

Sample ID

Finite pores

Fractal pores

∅ totala (%)

𝐷s

𝐷m

∅a (%)

𝐷s

𝐷m

∅a (%)

QD1-L3

/

2.64±0.05

3.70±0.02

2.78±0.01

/

2.10±0.01

5.81±0.03

QD1-L4

2.85±0.31

/

1.64±0.01

/

2.89±0.01

3.25±0.02

5.28±0.05

WX2-8

/

2.68±0.04

0.35±0.00

2.42±0.00

/

2.07±0.02

2.35±0.02

WX2-33

/

2.34±0.01

0.95±0.01

2.55±0.00

/

2.13±0.02

3.03±0.02

WX2-49

/

1.94±0.06

3.29±0.02

/

2.93±0.01

4.03±0.02

7.43±0.03

WX2-54

/

2.26±0.06

6.36±0.03

/

2.84±0.02

8.51±0.04

16.40±0.07

336 337 338 339

aThe

340

3.3.3 Pore structure affected by shale properties

𝑄 range for porosity estimation is between ~1.5×10-3 Å-1 and ~0.5 Å-1. ∅ total and each pore structure were estimated from total scattering intensity and scattering intensity of each pore structure, respectively. The experimental scattering intensity is used to estimate ∅ total while the modeled scattering intensity is used to estimate ∅ for each pore structure.

341

With the exception of sample WX2-49 there may be a U-shaped correlation between the fractal

342

dimension and TOC for the finite pores (Fig. 5b), while D increases and then becomes generally

343

constant with increasing TOC for the fractal pores (Fig. 5b). The maximum fractal dimension,

344

including both mass and surface fractals, occurs at ~27 % clay content and ~4.2 % TOC as shown

345

in Figs. 5a and b. This implies that the maximum complexity of shale fractal pore structure could

346

require at specific clay and carbon contents. In contrast, there is no obvious correlation between

347

the fractal dimension of the finite pores and clay content, and at best a weak decreasing or U-

348

shaped one for the fractal pores (Fig. 5a).

349

Figs. 5c and d show the total porosity and the porosities of the pore structures estimated by

350

both SANS and helium porosimetry as a function of clay and TOC contents. Negligible 19 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 44

351

correlations between the concentrations of other minerals and porosity were found; they do not

352

seem to be controlling factors for methane adsorption in these shales [27]. However, Figs. 5c and

353

d show that, in general, the SANS total porosity, the porosity of each pore structure, and the helium

354

porosity increase with increasing TOC. There is no obvious or weak correlation for clay content.

355

The results show a relatively strong correlation for TOC (R2 > 0.77) as a function of porosity,

356

indicating that TOC is a controlling factor in shale porosity. The weak correlation between clay

357

and porosity may reflect variations in the types of clay minerals present, as different clay minerals

358

may contribute to porosity differently.

359

20 ACS Paragon Plus Environment

Page 21 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

360

Fig. 5 Correlation between pore structure and chemical compositions for the six Chinese marine shales: Fractal

361

dimensions vs. (a) clay and (b) TOC contents; Porosities vs. (c) clay and (d) TOC contents (Note: WX2-49 was treated

362

as an outlier in Fig. 5b)

363

3.4 Methane adsorption characteristics

364

3.4.1 Adsorption capacity determination

365

Fig. 6 shows the results of the methane adsorption isotherm measurements. Methane

366

adsorption exhibited a two-stage change with increasing pressure for all samples measured. At low

367

to moderate pressures (0 to 4 MPa), the adsorption capacity rapidly increased with increasing

368

pressure, as indicated by the slope of adsorption profile. When the adsorption pressure exceeded

369

4 MPa, however, the rate of sorption gradually decreased. However, although the overall sorption

370

patterns are similar, the methane adsorption capacity dramatically varies among the different

371

samples. In general, as shown in Fig. 7b, the maximum adsorption capacity for methane in these

372

shales increased as a function of TOC content.

373 374

Fig. 6 Isothermal methane adsorption profiles for the six Chinese marine shales

21 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 44

375

The Langmuir adsorption model was used to fit the experimental methane adsorption isotherms

376

[67]. The Langmuir equation is usually recommended for describing methane adsorption in shale

377

samples due to its practical advantages and simplicity, as it only contains two important adjustable

378

parameters 𝑉L and 𝑃L. The usual form of the Langmuir model is [50]:

379

𝑉 = 𝑃L + 𝑃

380

where 𝑃 is the experimental pressure, 𝑉 is the experimental adsorption capacity at each pressure,

381

𝑉L is the Langmuir volume showing the theoretical maximum adsorption capacity, and 𝑃L is the

382

Langmuir pressure representing the pressure at which half of theoretical maximum adsorption for

383

a given sample occurs. As can be seen in Fig. 6, this model fits the measured data (Table 3) quite

384

well. The Langmuir sorption capacity, 𝑉L, varies from 0.47 cm3/g to 3.62 cm3/g and the Langmuir

385

pressure, 𝑃L, ranges from 0.76 MPa to 3.30 MPa.

386

𝑉L𝑃

(3)

Table 3 The Langmuir volume and pressure for the six Chinese marine shales Formation Niutitang (Є1)

Longmaxi (S1)

387

Sample ID

𝑉L (cm3/g)

𝑃L (MPa)

R2

QD1-L3

1.55

2.62

0.974

QD1-L4

1.86

2.39

0.969

WX2-8

0.47

0.76

0.974

WX2-33

1.15

3.30

0.987

WX2-49

3.18

1.72

0.987

WX2-54

3.62

1.54

0.990

3.4.2 Adsorption capacity affected by shale properties and pore structure

388

Fig. 7 shows the correlations between the Langmuir constants and the clay and TOC contents

389

of the shales. While there is some scatter, the Langmuir volume generally decreases with

390

increasing clay content and increases with increasing TOC (Figs. 7a and b). This implies that clay

391

minerals have a smaller methane adsorption capacity than organic matter. The Langmuir pressure 22 ACS Paragon Plus Environment

Page 23 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

392

𝑃L, however, decreases with both increasing clay content and TOC (Figs. 7c and d). Note that

393

sample WX2-8 is an outlier in Fig. 7d, possibly because this sample has the lowest TOC content

394

(1.52 %), highest plagioclase (> 14 %) and clay mineral contents (> 40 %) the lowest total porosity

395

(2.35 %) of the shale samples. 𝑃L is the pressure at which half of the maximum adsorption volume

396

is achieved, describing the affinity of the adsorbed gas for the adsorbent [50, 68]. Thus, the results

397

indicate that the affinities of methane for clay and organic matter in shale may be similar, although

398

the correlations for clay are much poorer than those for TOC. Again, this could be due to variations

399

in adsorption capability among different clay minerals. Regardless, the strong correlations between

400

the Langmuir constants and TOC (R2 > 0.83) suggest TOC may be a controlling factor for methane

401

adsorption in shales.

23 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 44

402 403

Fig. 7 Correlation between Langmuir constants and chemical compositions for the six Chinese marine shales:

404

Langmuir volume vs. (a) clay and (b) TOC contents; Langmuir pressure vs. (c) clay and (d) TOC contents (Note:

405

WX2-8 was treated as an outlier in Fig. 7d)

406

In order to quantify the effects of pore structure on methane storage in shales, the correlations

407

between the Langmuir constants and the pore structure parameters derived from the SANS data

408

need to be investigated. Figs. 8a and c show that the correlations between fractal dimension and

409

the Langmuir constants for the pore structure of the finite pores are negligible. However, there

410

may be U-shaped correlations between fractal dimension and the Langmuir constants for the pore

24 ACS Paragon Plus Environment

Page 25 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

411

structure of the fractal pores when the Langmuir constants are treated as the independent variables,

412

with maximum values of about 2.5 ml/g for 𝑉L and 2 MPa for 𝑃L.

413

The correlations between the Langmuir parameters and porosity are, however, somewhat better.

414

The Langmuir volume, 𝑉L, the maximum adsorption capacity, increases with increasing porosity

415

(Fig. 8b). This is reasonable because both porosity (Fig. 5d) and 𝑉L (Fig. 7b) increase with

416

increasing TOC. However, the Langmuir pressure, 𝑃L, the affinity of methane molecules for the

417

shale, decreases with increasing porosities, with WX2-8 again being an outlier (Fig. 8d). This result

418

also seems reasonable as porosities (Fig. 5d) and 𝑃L (Fig. 7d) increase with increasing TOC. The

419

results indicate that shale porosity has the opposite effect on the 𝑉L and 𝑃L values for methane;

420

higher porosity yields higher 𝑉L but lower 𝑃L for shale.

25 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 44

421 422

Fig. 8 Correlation between Langmuir constants and pore structure for the six Chinese marine shales: Langmuir volume

423

vs. (a) fractal dimensions and (b) porosities; Langmuir pressure vs. (c) fractal dimensions and (d) porosities (Note:

424

WX2-8 was treated as an outlier in Fig. 8d)

425

3.5 Shale pore accessibility

426

The combination of helium porosimetry with the SANS data described above, was used to

427

investigate pore accessibility quantitatively. Table 4 shows the estimated fraction of accessible

428

pores, 𝜑 for each of the shale samples. The 𝜑 should be an approximate value since the pore range

429

for SANS is approximate between 0.5 nm and 167 nm, while the pore size range of open pores

430

measured by He porosimetry is larger than 0.2nm [47]. In most cases, the measured 𝜑 value is 26 ACS Paragon Plus Environment

Page 27 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

431

greater than 64 %. This indicates that the matrix pores of the tested shales are highly accessible.

432

The exception is sample WX2-33, which has the lowest total pore accessibility (𝜑=40.26 %) and

433

helium accessible porosity (1.22 %) as well as the second lowest total porosity (3.12 %). This may

434

be caused by the mineralogy of this sample. WX2-33 has the highest calcite content (13.71 %) of

435

all the samples investigated. Late calcite precipitation may have filled or sealed off many of the

436

pores. For the other samples, however, the pore accessibility results indicate that the transport

437

pathways for methane production and CO2 sequestration are, in fact, relatively open. This suggests

438

that, to the extent that these samples are representative of the two formations, after drilling and

439

hydraulic fracturing are completed a significant amount of natural gas should be extractable from

440

the Longmaxi and Niutitang formations, possibly making them essential Chinese shale gas

441

resources.

442

Table 4 Revisit clay and TOC contents, and pore accessibility 𝜑 obtained from SANS and helium porosities for the

443

six Chinese marine shales Formation Niutitang (Є1)

Longmaxi (S1)

444

aThe

Sample ID

Clay (wt.%)

TOC (wt.%)

SANS ∅ total (%)

Helium ∅ (%)

𝜑a (%)

QD1-L3

29.59

3.29

5.81

3.72

64.03

QD1-L4

27.08

4.32

5.28

3.61

68.37

WX2-8

40.38

1.52

2.35

2.01

85.53

WX2-33

21.54

2.05

3.03

1.22

40.26

WX2-49

27.58

4.25

7.43

4.83

65.01

WX2-54

20.98

6.33

16.40

10.76

65.61

𝜑 values were estimated based on SANS ∅ total and helium ∅.

27 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 44

445 446

Fig. 9 Correlation between the fraction of accessible pores and chemical compositions for the six Chinese marine

447

shales: Fraction of accessible pores vs. (a) clay and (b) TOC contents (Note: WX2-33 was treated as an outlier)

448

Fig. 9 shows the correlations between the accessible pore fraction and the clay/TOC contents.

449

In contrast to the porosity correlations shown in Fig. 5, the fraction of accessible pores, 𝜑, increases

450

with increasing clay content (Fig. 9a), and decreases with increasing TOC (Fig. 9b). As noted

451

above, however, sample WX2-33 is an outlier with a 𝜑 value smaller than 41 %. These correlations

452

indicate that the fraction of accessible porosity in clay minerals is greater than that in organic

453

matter. Since the percentage of TOC (< 6.4 % for the tested shales) is small compared to the

454

mineral content, 𝜑 remains greater than 64 % for all these shale samples except WX2-33. Again,

455

the high calcite content of that sample (13.71 %) could be the cause of the low accessible porosity

456

(1.22 %) and pore accessibility (40.26 %) for this sample.

457

Fig. 10 shows the correlations between accessible porosity, 𝜑, and the fractal dimensions and

458

total, fractal and finite porosities. There are negligible correlations between the fractal dimensions

459

and the pore accessibility. However, 𝜑 seems to be anti-correlated with each of the porosity (again

460

with WX2-33 as an outlier). This may be because most of the inaccessible pores are in the organic

461

matter, so that samples with high TOC yield low 𝜑 (Fig. 9b) as well as high ∅ (Fig. 5d). These

28 ACS Paragon Plus Environment

Page 29 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

462

results suggest TOC has a more complex structure than the mineral matrix, and that the gas

463

transport pathways are in the minerals rather than in the organic matter, which is where of the

464

methane formed.

465 466

Fig. 10 Correlation between pore accessibility and pore structure for the six Chinese marine shales: Pore accessibility

467

vs. (a) fractal dimensions and (b) porosities (Note: WX2-33 was treated as an outlier in Fig. 10b)

468

3.6 Possible micro-structural reasons for different gas prospectivity

469

The nanoscale accessible pore structure of marine shales is strongly influenced by regional

470

geological conditions. Slight changes in geochemical and mineralogical parameters may greatly

471

influence their nanoscale pore system [69]. In this study, the Lower Silurian Longmaxi Formation

472

and Lower Cambrian Niutitang Formation are lower Paleozoic, organic-rich shales that were

473

deposited in a similar sedimentary environment with similar complex mineral compositions. Both

474

formations have experienced a complex tectonic history involving significant burial, uplift, and

475

erosion, and both are primary targets for shale gas development in China [1].

476

There are, however, large differences in shale gas production between these two formations

477

[3, 7, 70]. Significant commercial gas production has been developed from the Longmaxi shales,

478

but no commercial gas has yet been obtained from the Niutitang shales. The main difference 29 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 44

479

between the Niutitang and Longmaxi shales is paleo-burial depth. The Longmaxi shale was buried

480

to between 3000-7000 m in the study area, whereas the Niutitang was buried to greater than 6000

481

m. The thermal maturity of both shales reached the over-mature stage, but that of the Niutitang

482

shales was significantly higher than that of the Longmaxi. At this thermal maturity some of the

483

organic matter in the Niutitang shales has been carbonized. During carbonization, a subset of the

484

OM pores may have been destroyed or collapsed, resulting in a decrease of pore accessibility. In

485

addition, because it was buried deeper, the Niutitang Formation experienced more severe

486

compaction, which may have resulted in a decrease in pore size and a change in pore morphology.

487

These differences may partially explain why both 𝐷m and 𝐷s for the Niutitang shales are relatively

488

larger than those of Longmaxi shales (Table 2).

489 490

Fig. 11 Images of diverse graptolites in Longmaxi shales: (a) Abundant graptolites observed in core samples (sample

491

WX2-54); (b) Flattened carbonaceous graptolite fossil observed by SEM (sample WX2-54); (c) Interconnected pore

492

system within graptolite-derived OM (sample WX2-54)

493

An addition whereas all six samples examined in this study can be described as black

494

carbonaceous shales, abundant and diverse graptolite fauna are observed in Longmaxi shales but

495

not in Niutitang (Figs. 11a and b). Numerous nanoscale pores are well developed within the

496

graptolite-derived organic matter [24, 70], which could comprise a complex pore system which

497

may improve the pore accessibility of the Longmaxi shales (Fig. 11c). Despite all of these observed

498

and field-related differences, however, the results of this study (Table 4) show no obvious 30 ACS Paragon Plus Environment

Page 31 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

499

difference in pore accessibility between Longmaxi and Niutitang shales. This may simply be a

500

factor of the limited dataset in this study, but may also suggest that other factors, such as having

501

reached a minimum maturity level, are controlling, and that the observed differences in recovery

502

are related to larger-scale structures. Thus, a more comprehensive comparison study of the pore

503

structure of the two shale formations, and a correlation of the results with regional patterns of gas

504

production should be considered.

505

3.7 The implication of pore accessibility on methane production in shale

506

Fig. 12a and b show that both the Langmuir volume and pressure decrease with increasing

507

accessible pore fraction (with WX2-33 again an outlier). The higher the pore accessibility the lower

508

the maximum methane adsorption capacity (Fig. 12a). As noted above, 𝑉L increases with

509

increasing TOC (Fig. 7b), and the fraction of accessible pores deceases with increasing TOC and

510

increases with increasing clay content (Fig. 9). Thus, pore accessibility could be an indicator to

511

main chemical properties and methane storage in shale. It means that shale sample with higher

512

pore accessibility will have lower TOC, higher clay and lower methane sorption capacity, vice

513

versa. Since methane adsorption capacity is controlled by micropores, the results indirectly

514

confirm the conclusion that most of the micropores are in the organic matter, which tends to be

515

disconnected in shales. In contrast, higher pore accessibility correlates with lower maximum

516

methane adsorption pressure (Fig. 12b). Shales with a higher affinity for methane will have a lower

517

pressure at half of the maximum adsorption capacity. Thus, pore accessibility could be an indicator

518

to methane transport in shale. Shale sample with a higher pore accessibility tend to have lower 𝑃L

519

(higher methane affinity) - easier pathways for methane adsorption, vice versa, although our

520

sampling size is limited. As 𝑃L decreases with increasing TOC and clay content (Figs. 7c and d),

31 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 44

521

the affinity of methane for organic matter, combined with the pore accessibility could dominate

522

the degree of difficulty for methane recovery from the shale nanopores.

523 524

Fig. 12 Correlation between Langmuir constants and pore accessibility for the six Chinese marine shales: (a) Langmuir

525

volume and (b) Langmuir pressure vs. pore accessibility; (c) A schematic of adsorption isotherm affected by the

526

percentage of pore accessibility for shale (Note: WX2-33 was treated as an outlier in Fig. 12a)

527

While it may be surprising to find a relationship between the percentage of accessible pores and

528

methane adsorption capacity in shale, and this relationship needs further testing, if verified it may

529

be useful as a guide to methane production in shale gas reservoirs. Fig. 12c schematically shows

530

the effect of shale pore accessibility on methane adsorption isotherms given the relationships

531

between pore accessibility and 𝑉L and 𝑃L described above. The inverse correlation between the

532

pore accessibility and maximum adsorption capacity suggests that when pore accessibility is low,

533

decreasing transport pathways, maximum methane adsorption 𝑉L, low increases due to a

32 ACS Paragon Plus Environment

Page 33 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

534

simultaneous increase in adsorption sites. However, the data also suggest that under these

535

conditions the affinity of methane gas for the available sites in shale pores decreases. That is, the

536

pressure at half of the maximum adsorption capacity 𝑃L, low increases, so the amount of gas that

537

can be desorbed at a given pressure decreases. The methane adsorption capacity therefore increases

538

gradually with increasing adsorption pressure (Gray line in Fig. 12c). This has a negative effect on

539

shale gas production even with the high total methane adsorption capacity, because relatively little

540

gas is desorbed, and therefore releasable, at a given pressure.

541

Samples with high pore accessibility have a relatively small maximum adsorption, i.e., low

542

values of 𝑉L, high, but with steep adsorption profiles at low pressure and a plateau at high pressure,

543

i.e., low values of 𝑃L, high (Red line in Fig. 12c). That is, relatively more gas can be released at

544

lower pressures, but the maximum amount of gas storable is lower. Therefore, there may be an

545

optimum value for shale pore accessibility that will maximize production of methane in shale

546

reservoirs, and analysis of these values for target shales may help to direct drilling programs. In

547

the short-term, our data suggest that drilling in shale reservoirs with relatively high pore

548

accessibility will allow a significant amount of gas to be quickly produced at a relatively low

549

threshold pressures. In the longer-term, however, gradual production of larger total amounts of

550

methane with decreasing borehole pressures could be achieved for low pore accessibility shales.

551

Thus, the hydraulic fracturing plan may be tailored to control the percentage of accessible pores

552

in shale for the production time and amount during shale gas recovery.

553

4. Conclusions

554

In this study, we have combined FE-SEM, SANS, He porosimetry and volumetric high-

555

pressure adsorption techniques to characterize the geochemical and mineralogical compositions,

33 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 44

556

pore structure and methane adsorption of six samples from two Chinese marine shale formations.

557

On the basis of these results, the following conclusions have been drawn:

558

(1) FE-SEM images show that OM pores are the most important pore type, dominating the pore

559

systems of these shales. Intra-particle pores, inter-particle pores and microfractures resulting

560

from sedimentary diagenesis occur between or within mineral grains. Meanwhile, nanoscale

561

pores and microfractures are also influenced by mechanical compaction since both Longmaxi

562

and Niutitang Formations experienced multi-phase tectonic movements.

563

(2) Application of the unified fit model with two pore structures (finite pores and fractal pores) to

564

the SANS data shows that both surface and mass fractals are present. Mass fractals with

565

dimensions ranging between 1.94 and 2.68 are identified for finite pores except for sample

566

QD1-L4 which has a surface fractal dimension of 2.85. For fractal pores, samples QD1-L4,

567

WX2-49 and WX2-54 have mass fractal dimensions ranging between 2.84 and 2.93, while

568

samples QD1-L3, WX2-8 and WX2-33 have a surface fractal dimension ranging between 2.42

569

and 2.78. Total porosity, estimated by the Porod Invariant method, ranges between 2.35 % and

570

16.40 %, while porosity of finite pores ranges between 0.35 % and 6.36 %, and porosity of

571

fractal pores ranges between 2.07 % and 8.51 %.

572

(3) Pore accessibility estimated based on combined helium porosimetry and SANS shows that the

573

fraction of accessible pores is greater than 64 % for most of the samples, except for sample

574

WX2-33 which has only 40.26 % accessible pores, possibly as a result of carbonate

575

precipitation during diagenesis.

576

(4) Clay and TOC have opposite effects on the pore structure. Samples with higher clay contents

577

have higher pore accessibilities, and lower porosities and total methane adsorption capacities.

578

Increasing TOC has the opposite effect. This suggests that, in shales, gas is stored in the organic

34 ACS Paragon Plus Environment

Page 35 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

579

matter, but the transport conduits are through the inorganic matrix, especially through the clay.

580

Thus, the organic matter is the controlling factor for methane storage in shale gas reservoirs,

581

but the inorganic matrix controls gas release.

582

(5) In comparing the two shale formations, both the mass and surface fractal dimensions for

583

Niutitang shales are relatively greater than those of the Longmaxi shales. This may be a result

584

of the relative depths to which the two formations were buried (the depths from which the

585

Niutitang shales were recovered are greater than those of Longmaxi shales).

586

(6) Both Langmuir adsorption capacity and Langmuir pressure decrease with increasing shale pore

587

accessibility. This suggests that, in shales, high pore accessibility may reduce methane

588

adsorption capacity, while low pore accessibility may reduce methane desorption. Pore

589

accessibility could, therefore, become an essential factor in understanding and controlling

590

methane production in shale gas reservoirs.

591

Acknowledgements

592

The authors would like to acknowledge the financial support of the National Natural Science

593

Foundation of China (41802183), the National Postdoctoral Program for Innovative Talents

594

(BX201700282) and the China Postdoctoral Science Foundation (2017M621870). Work by LMA

595

was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy

596

Sciences, Chemical Sciences, Geosciences, and Biosciences Division. This research used

597

resources at the High Flux Isotope Reactor, a DOE Office of Science User Facility operated by the

598

Oak Ridge National Laboratory. Access to NG7 SANS was provided by the Center for High-

599

Resolution Neutron Scattering, a partnership between the National Institute of Standards and

600

Technology and the National Science Foundation under Agreement No. DMR-1508249. The

601

authors declare no competing financial interest. 35 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 44

602

Disclaimer

603

Certain commercial equipment, instruments, or materials are identified in this paper to foster

604

understanding. Such identification does not imply recommendation or endorsement by the

605

National Institute of Standards and Technology, nor does it imply that the materials or equipment

606

identified are necessarily the best available for the purpose.

607

References

608

[1] Zou, C.; Dong, D.; Wang, Y.; Li, X.; Huang, J.; Wang, S.; Guan, Q.; Zhang, C.; Wang, H.;

609

Liu, H.; Bai, W.; Liang, F.; Lin, W.; Zhao, Q.; Liu, D.; Yang, Z.; Liang, P.; Sun, S.; Qiu, Z. Shale

610

gas in China: Characteristics, challenges and prospects (I). Petrol. Explor. Dev. 2015, 42(6), 753-

611

67.

612

[2] Yang, F.; Ning, Z.; Wang, Q.; Zhang, R.; Krooss, B. M. Pore structure characteristics of lower

613

Silurian shales in the southern Sichuan Basin, China: Insights to pore development and gas storage

614

mechanism. Int. J. Coal Geol. 2016, 156, 12-24.

615

[3] Zhao, W.; Li, J.; Yang, T.; Wang, S.; Huang, J. Geological difference and its significance of

616

marine shale gases in South China. Petrol. Explor. Dev. 2016, 43(4), 547-59.

617

[4] Chalmers, G. R.; Bustin, R. M.; Power, I. M. Characterization of gas shale pore systems by

618

porosimetry,

619

microscopy/transmission electron microscopy image analyses: Examples from the Barnett,

620

Woodford, Haynesville, Marcellus, and Doig units. AAPG bull. 2012, 96(6), 1099-119.

621

[5] Javadpour, F.; Moravvej Farshi, M.; Amrein, M. Atomic-force microscopy: a new tool for gas-

622

shale characterization. J. Can. Petrol. Technol. 2012, 51(04), 236-243.

pycnometry,

surface

area,

and

field

emission

scanning

electron

36 ACS Paragon Plus Environment

Page 37 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

623

[6] Loucks, R.G.; Reed, R.M.; Ruppel, S.C.; Hammes U. Spectrum of pore types and networks in

624

mudrocks and a descriptive classification for matrix-related mudrock pores. AAPG bull. 2012,

625

96(6), 1071-98.

626

[7] Yang, C.; Zhang, J.; Tang, X.; Ding, J.; Zhao, Q.; Dang, W.; Chen, H.; Su, Yang.; Li, B., Lu,

627

D. Comparative study on micro-pore structure of marine, terrestrial, and transitional shales in key

628

areas, China. Int. J. Coal Geol. 2017, 171, 76-92.

629

[8] Ross, D. J.; Bustin, R. M. The importance of shale composition and pore structure upon gas

630

storage potential of shale gas reservoirs. Mar. Petrol. Geol. 2009, 26(6), 916-27.

631

[9] Loucks, R. G.; Reed, R. M.; Ruppel, S. C.; Jarvie, D. M. Morphology, genesis, and distribution

632

of nanometer-scale pores in siliceous mudstones of the Mississippian Barnett Shale. J. Sediment.

633

Res. 2009, 79(12), 848-61.

634

[10] Mastalerz, M.; He. L.; Melnichenko, Y.B.; Rupp, J. A. Porosity of Coal and Shale: Insights

635

from Gas Adsorption and SANS/USANS Techniques. Energy Fuels. 2012, 26(8), 5109-20.

636

[11] Clarkson, C. R.; Solano, N.; Bustin, R. M.; Bustin, A. M. M.; Chalmers, G. R. L.; He, L.;

637

Melnichenko, Y. B.; Radliński, A. P.; Blach, T. P. Pore structure characterization of North

638

American shale gas reservoirs using USANS/SANS, gas adsorption, and mercury intrusion. Fuel.

639

2013, 103, 606-616.

640

[12] Milliken, K. L.; Rudnicki, M.; Awwiller, D. N.; Zhang, T. Organic matter–hosted pore system,

641

Marcellus formation (Devonian), Pennsylvania. AAPG bull. 2013, 97(2), 177-200.

642

[13] Hao, F.; Zou, H.; Lu, Y. Mechanisms of shale gas storage: Implications for shale gas

643

exploration in ChinaMechanisms of Shale Gas Storage. AAPG bull. 2013, 97(8), 1325-1346.

37 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 44

644

[14] Chen, L.; Jiang, Z.; Liu, K.; Tan, J.; Gao, F.; Wang, P. Pore structure characterization for

645

organic-rich Lower Silurian shale in the Upper Yangtze Platform, South China: A possible

646

mechanism for pore development. J. Nat. Gas Sci. Eng. 2017, 46, 1-15.

647

[15] Sun, M.; Yu, B.; Hu, Q.; Chen, S.; Xia, W.; Ye, R. Nanoscale pore characteristics of the

648

Lower Cambrian Niutitang Formation Shale: A case study from Well Yuke #1 in the Southeast of

649

Chongqing, China. Int. J. Coal Geol. 2016, 154-155, 16-29.

650

[16] Pashin, J. C.; Kopaska-Merkel, D. C.; Arnold, A. C.; McIntyre, M. R.; Thomas, W. A.

651

Gigantic, gaseous mushwads in Cambrian shale: Conasauga Formation, southern Appalachians,

652

USA. Int. J. Coal Geol. 2012, 103, 70-91.

653

[17] Tian, H.; Pan, L.; Zhang, T.; Xiao, X.; Meng, Z.; Huang, B. Pore characterization of organic-

654

rich Lower Cambrian shales in Qiannan Depression of Guizhou Province, Southwestern China.

655

Mar. Petrol. Geol. 2015, 62, 28-43.

656

[18] Curtis, M. E.; Cardott, B. J.; Sondergeld, C. H.; Rai, C. S. Development of organic porosity

657

in the Woodford Shale with increasing thermal maturity. Int. J. Coal Geol. 2012, 103, 26-31.

658

[19] Kuila, U.; McCarty, D. K.; Derkowski, A.; Fischer, T. B.; Topor, T.; Prasad, M. Nano-scale

659

texture and porosity of organic matter and clay minerals in organic-rich mudrocks. Fuel. 2014,

660

135, 359-73.

661

[20] Chen, L.; Jiang, Z.; Liu, K.; Wang, P.; Gao, F.; Hu, T. Application of low-pressure gas

662

adsorption to nanopore structure characterisation of organic-rich lower Cambrian shale in the

663

Upper Yangtze Platform, South China. Aust. J. Earth Sci. 2017, 64(5), 653-665.

664

[21] Bernard, S.; Wirth, R.; Schreiber, A.; Schulz, H-M.; Horsfield, B. Formation of nanoporous

665

pyrobitumen residues during maturation of the Barnett Shale (Fort Worth Basin). Int. J. Coal Geol.

666

2012, 103, 3-11.

38 ACS Paragon Plus Environment

Page 39 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

667

[22] Houben, M. E.; Desbois, G.; Urai, J. L. Pore morphology and distribution in the Shaly facies

668

of Opalinus Clay (Mont Terri, Switzerland): Insights from representative 2D BIB–SEM

669

investigations on mm to nm scale. Appl. Clay Sci. 2013, 71, 82-97.

670

[23] Tian, H.; Pan, L.; Xiao, X.; Wilkins, R. W. T.; Meng, Z.; Huang, B. A preliminary study on

671

the pore characterization of Lower Silurian black shales in the Chuandong Thrust Fold Belt,

672

southwestern China using low pressure N2 adsorption and FE-SEM methods. Mar. Petrol. Geol.

673

2013, 48, 8-19.

674

[24] Ma, Y.; Zhong, N.; Cheng, L.; Pan, Z.; Dai, N.; Zhang, Y.; Yang, L. Pore structure of the

675

graptolite-derived OM in the Longmaxi Shale, southeastern Upper Yangtze Region, China. Mar.

676

Petrol. Geol. 2016, 72, 1-11.

677

[25] Li, J.; Zhou, S.; Li, Y.; Ma, Y.; Yang, Y.; Li, C. Effect of organic matter on pore structure of

678

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

679

China. Fuel. 2016, 185, 421-431.

680

[26] Schmitt, M.; Fernandes, C. P.; da Cunha Neto, J. A. B.; Wolf, F. G.; dos Santos, V. S. S.

681

Characterization of pore systems in seal rocks using Nitrogen Gas Adsorption combined with

682

Mercury Injection Capillary Pressure techniques. Mar. Petrol. Geol. 2013, 39(1), 138-49.

683

[27] Wang, Y.; Zhu, Y.; Liu, S.; Zhang R. Pore characterization and its impact on methane

684

adsorption capacity for organic-rich marine shales. Fuel. 2016, 181, 227-37.

685

[28] Anovitz, L. M.; Lynn, G. W.; Cole, D. R.; Rother, G.; Allard, L. F.; Hamilton, W. A.; Porcar,

686

L.; Kim, M. A new approach to quantification of metamorphism using ultra-small and small angle

687

neutron scattering. Geochim. Cosmochim. Ac. 2009, 73(24), 7303-7324.

688

[29] Anovitz, L. M.; Cole, D. R.; Rother, G.; Allard, L. F.; Jackson, A. J.; Littrell, K. C. Diagenetic

689

changes in macro- to nano-scale porosity in the St. Peter Sandstone: An (ultra) small angle neutron

39 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 44

690

scattering and backscattered electron imaging analysis. Geochim. Cosmochim. Ac. 2013, 102, 280-

691

305.

692

[30] Anovitz, L. M.; Cole, D. R.; Jackson, A. J.; Rother, G.; Littrell, K. C.; Allard, L. F.; Pollington,

693

A. D.; Wesolowski, D. J. Effect of quartz overgrowth precipitation on the multiscale porosity of

694

sandstone: A (U)SANS and imaging analysis. Geochim. Cosmochim. Ac. 2015, 158, 199-222.

695

[31] Anovitz, L. M.; Cole, D. R.; Sheets, J. M.; Swift, A.; Elston, H. W.; Welch, S.; Chipera, S.

696

J.; Littrell, K. C.; Mildner, D. F. R.; Wasbrough, M. J. Effects of maturation on multiscale

697

(nanometer to millimeter) porosity in the Eagle Ford Shale. Interpretation-J Sub. 2015, 3(3),

698

SU59-SU70.

699

[32] Wang, H-W.; Anovitz, L. M.; Burg, A.; Cole, D. R.; Allard, L. F.; Jackson, A.J.; Stack, A.

700

G.; Rother, G. Multi-scale characterization of pore evolution in a combustion metamorphic

701

complex, Hatrurim basin, Israel: Combining (ultra) small-angle neutron scattering and image

702

analysis. Geochim. Cosmochim. Ac. 2013, 121, 339-62.

703

[33] Swift, A. M.; Anovitz, L. M.; Sheets, J. M.; Cole, D. R.; Welch, S. A.; Rother, G. Relationship

704

between mineralogy and porosity in seals relevant to geologic CO2 sequestration. Environ. Geosci.

705

2014, 21(2), 39-57.

706

[34] Bahadur, J.; Melnichenko, Y. B.; Mastalerz, M.; Furmann, A.; Clarkson, C. R. Hierarchical

707

Pore Morphology of Cretaceous Shale: A Small-Angle Neutron Scattering and Ultrasmall-Angle

708

Neutron Scattering Study. Energy Fuels. 2014, 28(10), 6336-6344.

709

[35] King, H. E.; Eberle, A. P. R.; Walters, C. C.; Kliewer, C. E.; Ertas, D.; Huynh, C. Pore

710

Architecture and Connectivity in Gas Shale. Energy Fuels. 2015, 29(3), 1375-1390.

711

[36] DiStefano, V.H.; McFarlane, J.; Anovitz, L. M.; Stack, A. G.; Gordon, A. D.; Littrell, K. C.;

712

Chipera, S. J.; Hunt, R. D.; Lewis Sr., S. A.; Hale, R. E.; Perfect, E. Extraction of organic

40 ACS Paragon Plus Environment

Page 41 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

713

compounds from representative shales and the effect on porosity. J. Nat. Gas Sci. Eng. 2016, 35,

714

646-660.

715

[37] Yang, R.; He, S.; Hu, Q.; Sun, M.; Hu, D.; Yi, J.; Applying SANS technique to characterize

716

nano-scale pore structure of Longmaxi shale, Sichuan Basin (China). Fuel. 2017, 197, 91-99.

717

[38] Sun, M.; Yu, B.; Hu, Q.; Zhang, Y.; Li, B.; Yang, R.; Melnichenko, Y. B.; Cheng, G. Pore

718

characteristics of Longmaxi shale gas reservoir in the Northwest of Guizhou, China: Investigations

719

using small-angle neutron scattering (SANS), helium pycnometry, and gas sorption isotherm. Int.

720

J. Coal Geol. 2017, 171, 61-68.

721

[39] Xiao, X.; Wei, Q.; Gai, H.; Li, T.; Wang, M.; Pan, L.; Chen, J.; Tian, H.; Main controlling

722

factors and enrichment area evaluation of shale gas of the Lower Paleozoic marine strata in south

723

China. Petrol. Sci. 2015, 12(4), 573-586.

724

[40] Guo, X.; Li, Y.; Liu, R.; Wang, Q. Characteristics and controlling factors of micropore

725

structures of the Longmaxi Shale in the Jiaoshiba area, Sichuan Basin. Nat. Gas Ind. B 2014, 1(2),

726

165-171.

727

[41] Guo, T. Discovery and characteristics of the Fuling shale gas field and its enlightenment and

728

thinking. Earth Sci. Front. 2016, 23(1), 29-43.

729

[42] Schoenherr, J.; Littke, R.; Urai, J. L.; Kukla, P. A.; Rawahi, Z. Polyphase thermal evolution

730

in the Infra-Cambrian Ara Group (South Oman Salt Basin) as deduced by maturity of solid

731

reservoir bitumen. Org. Geochem. 2007, 38(8), 1293-1318.

732

[43] Wignall, G. D.; Littrell, K. C.; Heller, W. T.; Melnichenko, Y. B.; Bailey, K. M.; Lynn, G.

733

W.; Myles, D. A.; Urban, V. S.; Buchanan, M. V.; Selby, D. L.; Butler, P. D. The 40 m general

734

purpose small-angle neutron scattering instrument at Oak Ridge National Laboratory. J. Appl.

735

Crystallogr. 2012, 45, 990-998.

41 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 42 of 44

736

[44] Radliński, A. P.; Boreham, C. J.; Lindner, P.; Randl, O.; Wignall, G. D.; Hinde, A.; Hope, J.

737

M. Small angle neutron scattering signature of oil generation in artificially and naturally matured

738

hydrocarbon source rocks. Org. Geochem. 2000, 31(1), 1-14.

739

[45] Wignall, G. D.; Bates, F. S. Absolute calibration of small-angle neutron scattering data. J.

740

Appl. Crystallogr. 1987, 20, 28-40.

741

[46] Kline, S. R. Reduction and analysis of SANS and USANS data using IGOR Pro. J. Appl.

742

Crystallogr. 2006, 39, 895-900.

743

[47] Anovitz, L. M.; Cole, D. R. Characterization and Analysis of Porosity and Pore Structures.

744

Rev. Mineral. and Geochem. 2015, 80(1), 61-164.

745

[48] Glinka, C. J.; Barker, J. G.; Hammouda, B.; Krueger, S.; Moyer, J. J.; Orts, W. J. The 30 m

746

small-angle neutron scattering instruments at the National Institute of Standards and Technology.

747

J. Appl. Crystallogr. 1998, 31, 430-445.

748

[49] Ji, L.; Zhang, T.; Milliken, K. L.; Qu, J.; Zhang, X. Experimental investigation of main

749

controls to methane adsorption in clay-rich rocks. Appl. Geochem. 2012, 27(12), 2533-2545.

750

[50] Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am.

751

Chem. Soc.1918, 40, 1361-1403.

752

[51] Slatt, R. M.; O'Brien, N. R. Pore types in the Barnett and Woodford gas shales: Contribution

753

to understanding gas storage and migration pathways in fine-grained rocks. AAPG bull. 2011,

754

95(12), 2017-2030.

755

[52] Wang, Y.; Zhu, Y.; Chen, S.; Li, W. Characteristics of the Nanoscale Pore Structure in

756

Northwestern Hunan Shale Gas Reservoirs Using Field Emission Scanning Electron Microscopy,

757

High-Pressure Mercury Intrusion, and Gas Adsorption. Energy Fuels. 2014, 28(2), 945-955.

758

[53] O’Brien, N. R. Fabric of kaolinite and illite floccules. Clay Clay Miner. 1971, 19(6), 353-359.

42 ACS Paragon Plus Environment

Page 43 of 44 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

759

[54] Bahadur, J.; Radliński, A. P.; Melnichenko, Y. B.; Mastalerz, M.; Schimmelrnannv, A. Small-

760

Angle and Ultrasmall-Angle Neutron Scattering (SANS/USANS) Study of New Albany Shale: A

761

Treatise on Microporosity. Energy Fuels. 2015, 29(2), 567-576.

762

[55] Eberle, A. P. R.; King, H. E.; Ravikovitch, P. I.; Walters, C. C.; Rother, G.; Wesolowski, D.

763

J. Direct Measure of the Dense Methane Phase in Gas Shale Organic Porosity by Neutron

764

Scattering. Energy Fuels. 2016, 30(11), 9022-9027.

765

[56] Gu, X.; Cole, D. R.; Rother, G.; Mildner, D. F. R.; Brantley, S. L. Pores in Marcellus Shale:

766

A Neutron Scattering and FIB-SEM Study. Energy Fuels. 2015, 29(3), 1295-1308.

767

[57] Gu, X.; Mildner, D. F. R.; Cole, D. R.; Rother, G.; Slingerland, R.; Brantley, S. L.

768

Quantification of Organic Porosity and Water Accessibility in Marcellus Shale Using Neutron

769

Scattering. Energy Fuels. 2016, 30(6), 4438-4449.

770

[58] Jin, L.; Mathur, R.; Rother, G.; Cole, D.; Bazilevskaya, E.; Williams, J.; Carone, A.; Brantley,

771

S. Evolution of porosity and geochemistry in Marcellus Formation black shale during weathering.

772

Chem. Geol. 2013, 356, 50-63.

773

[59] Jin, L.; Rother, G.; Cole, D. R.; Mildner, D. F. R.; Duffy, C. J.; Brantley, S. L.

774

Characterization of deep weathering and nanoporosity development in shale-A neutron study. Am.

775

Mineral. 2011, 96(4), 498-512.

776

[60] Ruppert, L. F.; Sakurovs, R.; Blach, T. P.; He, L. L.; Melnichenko, Y. B.; Mildner, D. F. R.;

777

Alcantar-Lopez, L. A USANS/SANS Study of the Accessibility of Pores in the Barnett Shale to

778

Methane and Water. Energy Fuels. 2013, 27(2), 772-779.

779

[61] Mares, T. E.; Radliński, A. P.; Moore, T. A.; Cookson, D.; Thiyagarajan, P.; Ilavsky, J.; Klepp,

780

J. Location and distribution of inorganic material in a low ash yield, subbituminous coal. Int. J.

781

Coal Geol. 2012, 94, 173-181.

43 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 44 of 44

782

[62] Melnichenko, Y.B.; He, L.L.; Sakurovs, R.; Kholodenko, A. L.; Blach, T.; Mastalerz, M.;

783

Radliński, A. P.; Cheng, G.; Mildner, D. F. R. Accessibility of pores in coal to methane and carbon

784

dioxide. Fuel. 2012, 91(1), 200-208.

785

[63] Beaucage, G. Approximations leading to a unified exponential power-law approach to small-

786

angle scattering. J. Appl. Crystallogr. 1995, 28, 717-728.

787

[64] Ilavsky, J.; Jemian, P. R. Irena: tool suite for modeling and analysis of small-angle scattering.

788

J. Appl. Crystallogr. 2009, 42, 347-353.

789

[65] Zhang, R.; Liu, S.; Wang, Y. Fractal evolution under in situ pressure and sorption conditions

790

for coal and shale. Sci. Rep. 2017, 7(1), 8971.

791

[66] Glatter, O.; Kratky, O. Small angle X-ray scattering. Academic Press. 1982.

792

[67] Li, J.; Zhou, S.; Gaus, G.; Li, Y.; Ma, Y.; Chen, K.; Zhang, Y. Characterization of methane

793

adsorption on shale and isolated kerogen from the Sichuan Basin under pressure up to 60 MPa:

794

Experimental results and geological implications. Int. J. Coal Geol. 2018, 189, 83-93.

795

[68] Zhang, R.; Liu , S. Experimental and theoretical characterization of methane and CO2 sorption

796

hysteresis in coals based on Langmuir desorption. Int. J. Coal Geol. 2017, 171, 49-60.

797

[69] Chen, L.; Jiang, Z.; Liu, K.; Wang, P.; Ji, W.; Gao, F.; Li, P.; Hu, T.; Zhang, B.; Huang, H.

798

Effect of lithofacies on gas storage capacity of marine and continental shales in the Sichuan Basin,

799

China. J. Nat. Gas Sci. Eng. 2016, 36, 773-785.

800

[70] Wang, Y.; Zhu, Y.; Liu, S.; Chen, S.; Zhang, R. Comparative study of nanoscale pore structure

801

of Lower Palaeozoic marine shales in the Middle‐Upper Yangtze area, China: Implications for

802

gas production potential. Geol. J. 2018, 53, 2413-2426.

803

44 ACS Paragon Plus Environment