1 Time-lapsed visualization and characterization of shale diffusion

Accurate 3D-3D image registration was accomplished using MANGO software package62, which provides. 159 voxel-by-voxel precision. Further details on sa...
1 downloads 12 Views 5MB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Article

Time-lapsed visualization and characterization of shale diffusion properties using 4D X-ray microcomputed tomography Yulai Zhang, Peyman Mostaghimi, Andrew Fogden, Adrian Sheppard, Alessio Arena, Jill Middleton, and Ryan Troy Armstrong Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.7b03191 • Publication Date (Web): 02 Jan 2018 Downloaded from http://pubs.acs.org on January 2, 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 free 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 accessible to all readers and 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.

Energy & Fuels 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 25 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

Time-lapsed visualization and characterization of shale diffusion properties using 4D X-ray

2

microcomputed tomography

3

Yulai Zhanga, Peyman Mostaghimia, Andrew Fogdenb, Adrian Sheppardc, Alessio Arenab, Jill Middletonc,

4

and Ryan T. Armstronga*

5 6

a

7

b

8

c

9

*Corresponding author: [email protected]

School of Petroleum Engineering, The University of New South Wales, NSW 2052, Australia; FEI Oil & Gas, Suite 102, Level 1, 73 Northbourne Avenue Canberra, ACT 2600 Australia;

Department of Applied Mathematics, The Australian National University, Acton ACT 2601, Australia

10

Abstract

11

Diffusion is an important mass transport mechanism in shale matrix which usually has pore sizes ranging

12

from molecular dimensions to micrometers. Better characterization of the diffusion properties is helpful

13

in understanding the multi-physical mass transport process in shale. We present a method for

14

measuring local effective diffusivity of shale core plugs using 4D X-ray micro-computed tomography

15

(micro-CT). Liquid-liquid diffusion of X-ray opaque diiodomethane CH2I2 from a Permian Basin shale core

16

plug into the surrounding X-ray transparent toluene is monitored by 4D micro-CT imaging. The time-

17

sequenced diffusion tomograms enable 4D visualization of the dynamic process. Local directional

18

effective diffusivities are measured numerically from the micro-CT data using a mathematical method.

19

The measured data are analysed with relation to compositional variations of the sample. Dykstra

20

Parsons coefficient is used to quantify the degree of heterogeneity of the measured data at the sub-core

21

scale. We find that the diffusion in the Permian Basin sub-plug is uneven and influenced by matrix

22

heterogeneities. Dense materials, e.g. pyrite, have low porosity and low horizontal effective diffusivity of

23

around 10-15 m2/s or below; light materials, e.g. fossil, have high porosity and high horizontal effective

24

diffusivity of around 10-14 m2/s or above. Compositional variation of the sample leads to porosity and

25

mass transport property changes. 4D imaging and local diffusivity measurements identify the true

26

heterogeneity of the shale sample, which is advantageous over static imaging. The measured local

27

effective diffusivity enables us to infer smaller scale characteristics and thus provides a means to relate

28

microscale shale rock structure to macroscale transport properties.

29

1. Introduction

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 25

30

Shale gas has significantly changed the United States energy sector and also has great potentials in many

31

other countries around the world1-4. Although advanced drilling techniques and large scale hydraulic

32

fracturing have significantly increased gas production rates from low permeable shale rocks, our

33

fundamental knowledge of shale rocks remains far from complete1, 5, 6. Compared with conventional

34

sandstones, shale rocks are more heterogeneous in composition with complex pore space, e.g. shale

35

pores may range from molecular dimensions to micrometres, varying over several orders of magnitude7-

36

13

37

transport properties of shale is needed. As diffusion is a major mass transfer mechanism in the matrix

38

where the pore sizes are usually at nanometer length scale8, 14, diffusion studies in shale matrix are of

39

practical importance. The main purpose of this paper is to study the diffusion properties of shale rocks

40

within the shale matrix by using time-lapsed X-ray microcomputed tomography (micro-CT). The method

41

and validation for measuring effective diffusion coefficients from micro-CT experiments is published

42

elsewhere15. In this paper, we focus on the characterization of the shale sample by applying the time-

43

lapsed methodology to relate effective diffusivity values to shale properties and heterogeneity.

44

Molecular diffusion is the movement of molecules without bulk motion, and it is a result of the random

45

movement of molecules and molecule-molecule collisions according to Brownian motion. In a system

46

with different types of species, molecules tend to diffuse from higher concentration to lower

47

concentration, resulting in mixing of different substances, until an equilibrium state is reached.

48

Molecular diffusion is described by Fick’s law at the continuum scale, which states that diffusive flux is

49

proportional to the concentration gradient under the steady state condition and is defined as

50

 = − ∙ 

51

where  is diffusive flux, is the concentration of the diffusing substance and is the diffusion

. Better understandings on how the tremendous variation in length scales will affect the mass

(1)

52

coefficient. The theories and equations for the determination of for gases16, 17 and liquids are different

53

and since we are studying liquid-liquid diffusion in shale, only liquid diffusion fundamentals are

54

introduced here. For liquid diffusion, which is much slower than gas-gas diffusion, can be estimated

55

from the Stokes-Einstein equation, defined as

56

=

57

where  is the Boltzmann constant,  is solvent viscosity and  is radius of the diffusing particle. In eq 2,

58





(2)

does not depend on pressure, this is because liquids are only slightly compressible and pressure 2 ACS Paragon Plus Environment

Page 3 of 25 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

59

changes do not have a significant impact on the rate of collisions between molecules. In a binary system

60

with two types of species, may not be a constant and changes with the concentration of the diffusing

61

substance18, 19. The Stokes-Einstein equation (eq 2) is only applied to the diffusion of pure fluids. For

62

diffusion in porous media, due to reduced void space and torturous pore paths, the measured

63 64

becomes smaller and thus an effective diffusion coefficient  is commonly used to incorporate the

confining effect of a porous media. The ratio between  and , which is known as diffusive

65

conductance R, is related to accessible porosity , tortuosity , and constrictivity  20-23. A common

66

relationship for diffusive conductance is

67

=

68

which attempts to relate the macroscopic properties of a porous media to the confining effect that is

69

microscopic in nature. From the literature, R ranges from 0.05 to 0.28 for liquid diffusion in

70

unconsolidated porous media with porosity ranging from 0.1 to 0.3920. For gas diffusion in shale, R

71

values ranging from 0.013 to 0.038 were calculated using the lattice Boltzmann method (LBM) for

72

volumes constructed from scanning electron microscopy (SEM) images14. However, these volumes were

73

tiny, e.g. 500nm×500nm×1250nm, so the porosity (from 0.18 to 0.27) may not be representative. More

74

direct approaches use the method of volume averaging to consider the microstructure for determining

75

 in porous medium24-26. Overall, eq 3 is a general relationship that can be used to provide insight into

76

the microscale structure from macroscale measurements.

77

For diffusion in shale rocks, the previous researches can be roughly divided into two categories:

78

 



= 

(3)

(1) Building mathematical models that incorporate multiple physics, e.g. Darcy flow, slip flow,

79

Knudsen diffusion, adsorption and desorption, and then running simulations to determine the

80

relative importance of each mechanism in terms of their contribution to total gas flow.

81

(2) Measuring the diffusion coefficient of the porous media for a particular pair of fluids, typically

82

using adsorption/desorption isotherm studies for gas diffusion, and contact-mixing experiment

83

of two solutions for liquid diffusion.

84

In Category 1, Mehmani et al.27 considered both Knudsen diffusion and slip flow for each pore throat in a

85

pore network model and found that the apparent permeability of the pore network was significantly

86

increased at lower pressures. Huang et al.28 also took the pore network modelling approach and treated

87

gas conductivity enhancement by adding the Klinkenberg correlation. Chen et al.14 employed the LBM to 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

88

run nanoscale simulations on reconstructed pore structures of shale, from which effective Knudsen

89

diffusion coefficients were predicted. In Category 2, adsorption/desorption isotherm experiments are

90

frequently used to measure the diffusion coefficient of gases, typically CH4 or CO2, in crushed shale

91

samples29, 30. For this method, either a unipore model31 or bidisperse model32, 33 is used to fit the

92

experimental data, with the latter being capable of measuring two diffusion coefficients, namely macro-

93

pore and micro-pore diffusion coefficients. These methods are also routinely used for coals34, 35. Another

94

method is X-ray micro-computed tomography (micro-CT) imaging, which has been used extensively to

95

characterize porous rocks36-40 , study morphology and topology41, 42, model reactive transport43, 44, and

96

visualize multiphase flow45-49. Reviews on micro-CT techniques and their applications on porous

97

medium can be found in the works of Wildenschild and Sheppard50, Mostaghimi et al.51, Bultreys et al.52

98

and Blunt et al.53.

99

For studies of diffusion in porous medium using micro-CT, only a few papers can be found in the

100

literature. Guerrero-Aconcha and Kantzas54 measured the diffusion coefficient of propane in heavy oil

101

using computed assisted tomography. Liu et al. measured the local diffusion coefficients along the 1D

102

diffusion path of CO2 in n-decane saturated porous media55. Also, Tidwell et al. investigated the effects

103

of heterogeneous porosity on liquid diffusion in dolomite matrix using X-ray absorption imaging56. Polak

104

et al. studied the vertical diffusion of a liquid tracer from a fracture into the surrounding matrix of a

105

fractured chalk using a CT scanner and estimated the diffusion coefficient57. Vega et al. monitored the

106

diffusion of krypton gas in a coal sample using micro-CT58. However, in these studies the problems were

107

often simplified to 1D profiles for selected time steps. Although Agbogun et al. measured the diffusion-

108

accessible porosity of a dolostone sample and calculated temporal tracer concentrations in 3D during

109

diffusion, no diffusivity was measured59. Zhang et al.15, 60 measured the local directional effective

110

diffusivity of CH2I2 in toluene in shale plugs using 4D imaging using a new mathematical method. The

111

method treats each voxel as a continuum and is able to calculate local effective diffusivity by calculating

112

the local diffusive flux and local concentration gradient. However, the measurement was carried out

113

only on 2D domains, which would have a degree of uncertainty due to diffusion in the 3rd dimension.

114

Moreover, as the studied 2D domains were limited in size, the measured diffusivity data may not be

115

abundant enough to provide a statistical representation of the rock the sample. Therefore, full 4D

116

analysis of diffusion in shale is needed, which allows for studying the time evolution of the process and

117

directional anisotropy and thus, provides more reliable and abundant data. In addition, the link between

118

effective diffusion coefficient and rock heterogeneity also needs to be addressed.

4 ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25 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

119

Herein, we present 4D experimental data that provides an unprecedented wealth of information on

120

diffusion in shale rock. Similar to the experiment in Zhang et al.15, we use 4D micro-CT imaging to study

121

the liquid-liquid diffusion of diiodomethane CH2I2 from a Permian Basin shale core plug surrounded by

122

toluene. The spatial and temporal progress of shale diffusion is visualized in sequenced 3D tomograms –

123

4D imaging. The novelty of this work is that we are able to quantify diffusion coefficients at the

124

micrometer length scale while visualizing larger more representative 3D samples rather than previous

125

works that are only 1D or 2D. Imaging techniques commonly suffer from the trade-off between field of

126

view and image resolution7. The technique proposed herein bridges this gap and thus provides a means

127

to relate shale rock structure to macroscale transport properties. We quantify the variability of diffusion

128

coefficients within a shale core plug, and then we link the observed macroscopic phenomenon to

129

microscopic pore structures and heterogeneities. Moreover, by measuring the local effective diffusivity

130

of the whole plug in 3D, we are able to make more reliable statistical characterization and quantification

131

of sample anisotropy. The presented technique is not limited to shale rock and could be applied to any

132

porous system and in particular highly heterogeneous materials. Our work provides a mean to develop

133

large scale models that directly link to the underlying structure and sub-scale heterogeneity of porous

134

systems.

135

2. Materials and Methods

136

A shale sub-plug from the Permian Basin formation was used. The sub-plug was drilled perpendicular to

137

the bedding plane out of a centimeter-sized core plug. The sub-plug was 4.3mm in diameter and 6.2mm

138

in height. Bulk analysis of sister sample material gave a porosity of 7.8% from Helium porosimetry and a

139

total organic carbon (TOC) content of 3.0 wt% from the LECO method. The thermal maturity of this

140

siliceous shale lies in the early-mid oil window, and its mineral composition from XRD is listed in Table 1.

141

The sub-plug was scanned over its full height, at a voxel size of 1.9 µm, using a double-helically

142

trajectory on a HeliScan micro-CT scanner61 in a sequence of 2 states, namely 1) after cleaning and

143

drying and 2) after saturation with CH2I2. Cleaning was performed using toluene and methanol to

144

remove hydrocarbons, soluble bitumen, water and salts. Saturation of the cleaned and dried sample

145

with CH2I2 was performed in an off-line process using vacuum infiltration followed by isostatic

146

pressurization to 8000 psi for 10 days. In particular, CH2I2 is used as a contrast agent due to its high

147

opacity; however, this also tends to produce imaging artefacts such as beam hardening, which must be

148

dealt with during image processing.

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 25

149

A series of time-sequenced tomograms during diffusion were acquired by continuous circular scanning

150

(at lower resolution, 5 μm/voxel) of a vertical sub-section of the initially CH2I2-saturated sub-plug after

151

its immersion in a toluene-filled holder under ambient conditions. This series of micro-CT images

152

monitors the diffusion of CH2I2 from the sub-plug into the surrounding toluene. Table 2 lists the duration

153

of each circular scan, number of circular scans, image size, resolution of tomograms, and the field of

154

view for the experiment. All tomograms were corrected for beam hardening, masked, spatially aligned,

155

and linearly rescaled such that the grayscale values (attenuation) of solid grains matched across all

156

imaged states. Beam hardening correction (BHC) was performed on the dry-state and saturated-state

157

tomograms and the first tomogram in the diffusion series, to flatten their radial profiles, and the same

158

correction to this first diffusion tomogram was applied to all subsequent tomograms in the series.

159

Accurate 3D-3D image registration was accomplished using MANGO software package62, which provides

160

voxel-by-voxel precision. Further details on sample preparation and image analysis are provided in a

161

previous conference paper by Zhang et al. (2017)57.

162

Table 1. Sample characterization data with minerals presented as wt% of total mineral Porosity

TOC

Calcite

(%)

(wt%)

(wt%)

7.8

3.0

0

Illite/

Illite/

Quartz

Smectite

Muscovite

Plagioclase

(wt%)

(wt%)

(wt%)

5

25

65

Pyrite (wt%) 5

163 164

Table 2. Micro-CT acquisition parameters Duration of each circular scan (min) 36

Number of circular scans 60

Height of field of view (mm) 2.6

Tomogram size

Resolution

(voxels)

(µm/voxel)

871×865×520

5.0

165 166 167

The mathematical method for measuring local effective diffusivity values is presented and validated in

168

our previous work.15. Herein, we extend our previous work by implementation of the conjugate gradient

169

methods to solve the system of equations for a large 3D domain. This allows for quantification of vertical

170

and horizontal effective diffusivities, which were not possible in our previous validation work.13

171

Therefore, only a brief introduction of the method is provided below.

6 ACS Paragon Plus Environment

Page 7 of 25 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

172

Relative concentration and relative density fields of CH2I2 are calculated from the grayscale values of the

173

tomograms. We calculate relative mass concentration of CH2I2 ( ) as

174

(!,#,$,%) = &

175

where -./011(2,3,4,5) is the grayscale at location (6, 7, 8) of the diffusion tomogram at time t, and

& '())(!,#,$,%) *& '#(!,#,$)

(4)

&+,(!,#,$) *& '#(!,#,$)

176

-.9:;) is obtained from

179

?(!,#,$,%) =

180

where the constant -.AB;C; is the grayscale of a voxel in a resolved pore in the state of full saturation

& '())(!,#,$,%) *& '#(!,#,$)

(5)

& &+, *& @(

181

with CH2I2 and -.D0= is the corresponding value in the dry (air-filled) state. Relative density values ranges

182

between 0 and the local porosity of the voxel.

183

Fick’s second law, eq 6, is used to model diffusion. To measure the local effective diffusion coefficients,

184

the spatial variation of  is considered as

185

E?(!,#,$,%) E%

186

where  is the effective diffusion coefficient at each location and each has three components, i.e. H, I

187 188

= F ∙ ( F(!, #, $, %))

(6)

and J directions. Here we assume the spatial variation of  is only a result of pore geometry changes in

the space, i.e.  does not change with local compositional changes. So  is assumed to be constant for

189

each location and for each direction. To solve for  , a scalar potential field, P, is introduced as

190

FK =  = − F(!, #, $, %)

191

where J is diffusive flux. Then eq 6 becomes

192



193

where the left-hand side can be estimated by the temporal difference of two density fields using the

194

finite difference method, i.e. >5LM5 and >5 . eq 8, which is formally known as Poisson’s equation, leads to

195

a large system of linear equations after discretising using finite difference method (FDM) for the 3D

196

images (871×865×520 voxels) obtained. The conjugate gradient (CG) method is used to solve for the

E?(!,#,$,%) E%

(7)

= F K(!, #, $, %)

(8)

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 25

197

solutions, which takes around 3 hours on 8 CPU cores and 100GB of memory. After solving for the

198

potential field, P, the effective diffusion coefficient at each voxel in each direction can be calculated by

199

( )( = − (F)( , ( = N, O, P

200

where i signifies all 3 dimensions. However, the local diffusion coefficient results are more suitably

201

presented in terms of horizontal local effective diffusion coefficient Q and vertical local effective

(FK)

(9)

(

202

diffusion coefficient R . R is the J direction local diffusion coefficient. Q is in the HI plane, parallel to

203

bedding, and calculated by

204

S(!,#,$) =

205

where N(!,#,$) and O(!,#,$) are the diffusive flux at location (6, 7, 8) in H and I direction, respectively.

TN(!,#,$)  LO(!,#,$) 

(10)

TFN(!,#,$)  LFO(!,#,$) 

206

Also, FN(!,#,$) and FO(!,#,$) are the concentration gradient at location (6, 7, 8) in H and I direction,

207

respectively.

208

Although the pore spaces in shale are sub-resolution at 5 μm/voxel, the local porosities can be

209

estimated from tomograms by dividing density (>), eq 5, by concentration of CH2I2 (φ), eq 4, which

210

provides

211

(!,#,$)

212

where (2,3,4) is the local porosity of each voxel and does not change with time. Dykstra-Parsons (DP)

213

coefficients63 are calculated to quantify the variation of local parameters in 3D space by using

214

K =

215

where [̅ is the mean value of a given variable, which could be mean grey-scale, porosity or effective

=

& &+,(!,#,$) *& '#(!,#,$)

(11)

& &+, *& @(

V *'WX.Z ' V '

(12)

216

diffusion coefficient, i.e. a variable value with 50% probability. Also, []^._ is the value of the same

217

parameter with 84.1% probability, which is mean plus one standard deviation.

218

3. Results and Discussion

219

The images in Figure 1 show the same central longitudinal slice of the helically-scanned tomogram of the

220

Permian Basin sub-plug in its (a) dry and clean state, and (b) CH2I2 saturated state after registration. In 8 ACS Paragon Plus Environment

Page 9 of 25 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

221

Figure 1 (a), the prevalent bright features comprise dense mineral, which from Table 1 is pyrite. Aside

222

from a couple of visible fractures, which taper in from the sub-plug periphery and appear to be coring-

223

induced, the darkest features are chambers with slightly brighter interiors – presumably comprising low

224

density authigenically-formed mineral – which more commonly occur in the upper half. In some

225

instances, mainly in the lower half, this interior mineral phase has grown and densified to leave only a

226

thin dark ring around it. These dark regions, and those in thin, short seams, may comprise pore or

227

organic matter or a mixture of both of these low attenuating phases. The remainder of the sub-plug is a

228

fairly featureless, uniform grayscale in Figure 1(a), within which the voxels presumably contain a fine

229

mixture of mineral grains, clays, porosity and organic matter11.

230

After saturation with CH2I2, the brightening of voxels from Figure 1(a) to Figure 1(b) directly reflects the

231

local connected porosity at that location. All dark features in Figure 1(a) strongly brighten and thus

232

possess substantial porosity. The low density mineral interiors of chambers are also highly porous, while

233

the denser interiors naturally brighten to a lesser degree. The uniform grey background regions in Figure

234

1(a) also brighten due to matrix porosity, to reveal features and textures that were not apparent in the

235

dry state. In particular, the saturation exposes two distinct rock types in the upper and lower halves of

236

the sub-plug in Figure 1(b). The upper half has somewhat lower porosity, which is also more

237

homogeneously distributed. Occasional dark spots in the matrix correspond to solid grains of

238

quartz/plagioclase (from Table 1). Although these grains are much larger than the voxel size of 1.9 µm,

239

they cannot be distinguished in the dry state due to the similar attenuation of the surrounding matrix of

240

clays of illite and smectite (from Table 1), but become apparent in Figure 1(b) due to the brightening of

241

the latter. The porosity in the lower half is somewhat higher and is more heterogeneously distributed,

242

exhibiting darker sub-laminations of low porosity that generally coincide with pyrite-lean bands. Note

243

that the above-mentioned fractures from the sub-plug extremities are mainly not brightened since CH2I2

244

drained from these regions during acquisition of the saturated-state tomogram, driven by the large

245

density difference between CH2I2 and air. However, a number of even finer fractures within the sub-plug

246

interior become apparent due to the saturation treatment, as discussed further below.

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 25

1mm 247 248

(a)

(b)

249 250

Figure 1. Central longitudinal slice (4.3mm x 6.2mm) of the helically-scanned tomogram of the Permian Basin sub-plug (a) in its

251

circular scans during the subsequent diffusion experiment.

252

The yellow rectangle in Figure 1, straddling the two rock types, shows the reduced vertical field of view

253

of the circular scans of this sub-plug during the diffusion experiment. Figure 2(a) and 2(f) display this

254

same slice of these two tomograms after cropping to this reduced height. Figure 2 (b-e) show the

255

corresponding registered tomogram slice at four times during 4D imaging of the diffusion experiment,

256

namely at 2h 42m, 5h 42m, 17h 6m, and 33h 18m, respectively. These time-sequenced tomograms

257

display the progress of diffusion by the darkening of the images from Figure 2(a) to (e), due to the local

258

loss of attenuation arising from the surrounding toluene (dark) entering the sub-plug to mix with and

259

dilute the CH2I2 (bright).

260

At the very start of the diffusion series (e.g. Figure 2(b)), more fine fractures are brightened than in the

261

saturated state of Figure 2(a), since the 4D imaging (starting from the CH2I2 re-saturated state) involves

262

less fracture drainage owing to the reduced fluid density difference and the faster imaging. The diffuse

263

darkening of the fractures near the sub-plug periphery at these early stages is driven by fast diffusion of

clean & dry state and (b) after saturation with CH2I2 and registration. The yellow rectangle indicates the field of view of the

10 ACS Paragon Plus Environment

Page 11 of 25 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

264

toluene along these pathways, while radial ingress and mixing of toluene through the unfractured matrix

265

regions lags behind. Over time, the diffusive darkening advances and proceeds most rapidly where the

266

toluene meets up with fine interior fracture segments and/or locations where matrix porosity is higher.

267

The high porosity chambers only admit toluene once their surrounds do, and thus are not preferred

268

pathways. Diffusion is faster in the lower rock type; midway through the imaged experiment the

269

darkening has spanned this lower half while the upper half remains bright across its centre. Near the

270

end of the imaging series, the grayscales in the lower half in Figure 2(e) approach those in the dry state

271

of Figure 2(f), since toluene attenuates only slightly greater than air, while diffusion remains incomplete

272

in middle of the upper half, especially apparent from the chambers there in which their CH2I2 remains

273

substantially undiluted.

274 275

(a)

(b)

(c)

(d)

276 277

1mm 278 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

279

(e)

Page 12 of 25

(f)

280

Figure 2. Central longitudinal slice (4.3mm x 2.6mm) of the registered tomogram series of the Permian Basin sub-plug, in its (a)

281 282

CH2I2 saturated state and (f) clean & dry state, reduced from Figure 1 to the field of view of the diffusion experiment, and (b)-(e)

283

Figure 3 displays the same field of view of these same six states for the corresponding difference

284

tomograms, i.e. CH2I2-saturated state minus diffusing-state. This difference serves to subtract the

285

contributions from all solid components (minerals and organic matter) to isolate the changes in CH2I2

286

concentration in pores due to the dilution of CH2I2 by toluene. In particular, these images locally

287

brighten in proportion to the loss in volume of CH2I2 at each voxel. Near the start of the diffusion

288

imaging (e.g. Figure 3(b)), the difference is only bright where diffusion of toluene into the sub-plug has

289

commenced, primarily along and out from fractures. The remainder is dark, indicating no CH2I2 dilution

290

at this stage. The darkest fractures correspond, as mentioned above, to those that drained in the

291

saturated-state image and thus contain more CH2I2 at the start of the diffusion series. The brightening of

292

the difference tomograms continues from Figure 3(b) to 3(e), proceeding fastest along the more

293

preferred pathways, i.e. fractures. This last image is very similar in its lower half to the difference CH2I2-

294

saturated state minus dry state in Figure 3(f), while the dark cloud over the center of the upper half

295

demonstrates the incompleteness of diffusion there. The contribution of locally high porosity and/or

296

fractures to diffusion, based on the above-mentioned calculation of concentration and density from

297

such tomogram differences, is quantified below.

showing diffusion tomogram at time of 2h42m, 5h42m, 17h6m, and 33h18m, respectively.

298 299

(a)

(b)

12 ACS Paragon Plus Environment

Page 13 of 25 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

300 301

(c)

(d)

1mm 302 303

(e)

(f)

304 305

Figure 3. Central longitudinal slice (4.3mm x 2.6mm, with field of view corresponding to Figure 2) of the registered tomograms

306

saturated minus diffusing-state difference tomograms at time of 2h42m, 5h42m, 17h6m, and 33h18m, respectively.

307

We selected a late time diffusion tomogram, 30 hours after the start of the diffusion experiment, for full

308

3D simulation using eq 4 – eq 9. This particular time was selected to make sure that diffusion has

309

occurred in most of the voxels such that we have a concentration gradient across the sample that allows

310

for measurement of local diffusion coefficients. In Figure 4 (a) and (b) we display horizontal ( Q ) and

311

vertical ( R ) effective diffusion coefficients in 3D for the entire field of view. These results provide a

312

direct means to evaluate the spatial variation of effective diffusivities with relatively high resolution.

313

With red colors indicating higher values while blue colors indicating lower values, heterogeneities are

314

clear in both of the two volumes, which we will discuss in detail.

315

Special attention is required when interpret the measurement results because they may be affected by

316

the mathematical method and/or tomogram noise. Three types of data are thought to be unreliable

317

with our current methods and may lead to erroneous diffusivity measurements. The 1st type of error is

318

from data in the middle of the core plug, i.e. around the central axis. Within this region Q values are

319

of the Permian Basin sub-plug in (a) its dry state, (f) its CH2I2-saturated minus dry-state difference; (b)-(e) 4 slices of the CH2I2-

smaller compared with those in the outer regions. These smaller Q do not necessarily mean low 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

320

diffusivity, they are a result of the method. For our diffusion experiment, in which CH2I2 diffuses radially

321

from the sub-plug into the surrounding toluene, the central axis has a symmetrical no-flux boundary

322

condition. As a result, the diffusive flux and concentration gradient near the central axis are too small to

323

be accurately captured. The 2nd type of error is in the outer region along the perimeter of the R volume.

324

The obtained R values are obviously smaller than compared with those in the inner region, which is not

325

a true reflection of the diffusion properties of the sample. The reason for this is thought to be that

326

vertical diffusion in this region is too small to be accurately captured. For our experiment set-up, the

327

principal diffusion is in the radial direction and vertical diffusion is only a result of heterogeneous radial

328

diffusion. In other words, unless there is vertical concentration gradient caused by uneven horizontal

329

diffusion, there will not be vertical diffusion. Because the outer region is close to the boundary, diffusion

330

is more complete in all directions and like a steady state which leads to relatively homogenous

331

concentration field. As a result, vertical diffusion is small. By contrast, in the inner region which is far

332

from the boundary, diffusion is more controlled by fractures. Fractures create fast radial diffusion from

333

deep inside the sub-plug, which results in uneven horizontal diffusion and heterogeneous concentration

334

field. So, vertical concentration gradient and vertical diffusion in the inner region are larger in the inner

335

region. The 3rd type of error is evident as white regions in the Q and R volumes, which are too high

336

values, e.g. infinity. They are prone to occur in regions where almost no concentration gradient of CH2I2

337

exists and no diffusion occurs. These regions correspond to materials that have too low porosity, like

338

pyrite and quartz. Because their porosity is very low, the attenuation change due to CH2I2 diffusion is too

339

small and easily blurred by image noise. As a result, erroneous concentration gradient of CH2I2 is

340

calculated which leads to faulty diffusivity measurement. For the remainder of our analyses, we

341

disregard regions of the sample that have erroneous local diffusivity measurement.

342

In Figure 4 (c) and (d) we display the histograms of the effective diffusivity data in Figure 4 (a) and Figure

343

4 (b), respectively after removing data along the boundaries. That means only Q data that have a

344

horizontal distance to the central axis larger than 100 voxels are used for Figure 4 (c) and only the R

345

data that have a horizontal distance to the central axis smaller than 300 voxels are used for Figure 4 (d).

346

The histograms show that about 98% of Q values are between 10-15 and 10-13 m2/s while R values are

347

more distributed with about 80% in the range between 10-16 and 10-14 m2/s. Overall, the R is generally

348

lower than the Q , indicating anisotropic diffusion property, which corresponds to the sample being cut

349

parallel to the bedding plane.

14 ACS Paragon Plus Environment

Page 14 of 25

Page 15 of 25 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

1mm 350 351

(a)

(b)

(c)

(d)

352 353 354 355 356

Figure 4. Local effective diffusion coefficient calculation results in 3D of the full field of view of the diffusion experiment with (a) showing the S and (b) showing the ` . Each data volume has a size of 430×520 (radius × layer) voxels; (c) and (d) are the histograms of the S and ` volume, respectively.

357

Figure 5 and Figure 6 show (a) two cross-sections of the Q volume of the Permian Basin sub-plug, and

358

their corresponding (b) porosity maps, (c) dry tomograms, and (d) CH2I2 saturated tomograms. These

359

images are taken from a plane parallel to the bedding. In Figure 5 (a) the Q data appears relatively

360

homogeneous without too many abrupt variations. The dry and CH2I2 saturated tomograms, Figure 5 (c)

361

and (d), also show a relatively homogeneous cross-section in terms of attenuation, which correspond 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 25

362

well to the porosity map of Figure 5 (b). The vast gray background in the dry tomogram, Figure 5 (c),

363

corresponds to the shallow green areas in Figure 5 (b) that have porosities between 5% and 10% and the

364

continuous areas that are mostly red in Figure 5 (a), which have local effective diffusion coefficients of

365

around 10-14 m2/s. As discussed earlier, they presumably contain a fine mixture of mineral grains, clays,

366

porosity and organic matter. The bright spots in the dry tomogram, which are most likely dense pyrite

367

crystals, have lower porosities below 5% and also lower local effective diffusion coefficients of around

368

10-15 m2/s. They correspond to the small blue clusters in Figure 5 (a). The more porous (with porosities

369

above 20%) and less dense chambers, presumably comprising low density authigenically-formed mineral,

370

are very bright in the CH2I2 saturated tomogram with more CH2I2 admitted. They are dark red in the Q

371

map with high local effective diffusion coefficients above 10-14 m2/s. One common feature of these

372

heterogeneities is that they all have a relatively small correlation length. No clear evidence of fractures

373

is observed for this particular cross-section, which suggests a relatively homogeneous diffusion with only

374

small-scale heterogeneities that are a result of spatial compositional changes in the sub-plug. The DP

375

coefficient of the Q data in Figure 5 (a) is 0.57.

376 377

(a)

(b)

1mm

378 379

(c)

(d)

16 ACS Paragon Plus Environment

Page 17 of 25 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

380 381

Energy & Fuels

Figure 5. A cross-section of the Permian Basin sub-plug with (a) showing the measured local effective S in unit of m /s, (b) 2

showing the porosity map, (c) showing the dry & clean tomogram and (d) showing the CH2I2 saturated tomogram.

382

The images provided in Figure 6 demonstrate a region of the shale sample that is quite different from

383

the one in Figure 5. One most obvious example is the large blue regions that have lower local effective

384

diffusion coefficients in the upper part of Figure 6 (a). They correspond to the lower-porosity less-

385

permeable area in the porosity map, Figure 6 (b). These regions are also evident in the CH2I2 saturated

386

tomogram, Figure 6 (d). For shale rock, the pore space of these regions may have pore throats that are

387

only several nanometers in diameter and diffusive pathways may be highly torturous and not well

388

connected. Although the materials for these regions have similar density as other regions, judging from

389

the moderate and homogeneous attenuation values in the dry and clean tomogram, they have lower

390

porosity, which leads to slow mass transport. This change of composition is clearly captured by the Q

391

map and CH2I2 saturated tomograms. This once again suggests that the attenuation properties of the

392

clean and dry tomograms fail to identify heterogeneity. The importance of identifying these low porosity

393

materials is that these regions will have a negative effect both on gas storage and gas flow in a shale gas

394

reservoir, quantification of the weight or volumetric percentage of these materials will give us a rough

395

idea of the quality of the sample in terms of reservoir rock. It would also be helpful for the economic

396

evaluation of a shale gas play if a large enough number of samples can be analysed. It is interesting that

397

even in the low porosity region we can find some high porosity high permeable regions, i.e. the dark red

398

spots in the upper region of Figure 6 (a) which are yellow with porosities above 20% in Figure 6 (b). This

399

reflects the complex composition and structure of the sample. The DP coefficient of Q of the cross-

400

section in Figure 6 is 0.59, which is higher than the one presented in Figure 5, although the two cross-

401

section planes are only 0.34 mm apart.

402 403

(a)

(b)

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 25

1mm 404 405 406 407 408

(c)

(d)

Figure 6. A cross-section of the Permian Basin sub-plug, 0.34mm below the one shown in Figure 5, with (a) showing the measured local effective S in unit of m /s, (b) showing the porosity map, (c) showing the dry & clean tomogram and (d) 2

showing the CH2I2 saturated tomogram.

409

To study the variation of the degree of heterogeneity between cross-sections, DP coefficient of effective

410

horizontal diffusivity Q is calculated for each cross-section from top to bottom of the field of view, with

411

results shown in Figure 7 (a). The results show fairly evident changes in the DP coefficient for effective

412

diffusivity within the scanned section of the sub-plug, which is only 2.6 mm in length. The upper section

413

has clearly lower DP coefficients than the lower section, indicating that the lower section is more

414

heterogeneous. This corresponds well to the observations we have in the high-resolution saturated

415

tomogram, Figure 1 (b), in which the rock type in the lower half of the sub-plug is more heterogeneous.

416

These findings demonstrate the heterogeneity in different layers and may also be helpful in studying the

417

laminations of shale on a micrometer scale. For comparison studies, the DP coefficients of grayscale

418

values of the CH2I2 saturated and dry and cleaned tomograms are also calculated and shown in Figure 7

419

(b). However, quite different to Figure 7 (a), Figure 7 (b) demonstrates much more homogeneity since

420

the DP coefficients are generally much smaller and do not change much between cross-sections. So 4D

421

imaging and local diffusivity measurements identify the true heterogeneity of the shale sample.

18 ACS Paragon Plus Environment

Page 19 of 25

Dykstra Parsons coefficient 0.55

0.6

0.65

Dykstra Parsons coefficient

0.7

0

422 423 424 425 426

0.05

0.1

0.15

1 Cross-section number

1

Cross-section number

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

101 201 301

101 201 301

401

401

501

501 Effective diffusivity

Grayscale - wet tomo

(a)

Grayscale - dry tomo

(b)

Figure 7. DP coefficient of each cross-section of (a) the effective S volume, (b) the CH2I2 saturated and dry & clean tomogram of the Permian Basin sub-plug, from top to bottom.

In Figure 8, R maps of vertical central longitudinal slices of the Permian Basin sub-plug are cut from the

427

3D volume. The R maps in Figure 8 are more heterogeneous than the Q maps in Figure 5 and Figure 6,

428

with large blue regions having low effective diffusion coefficients. However, as discussed earlier, the

429

blue areas of the sub-plug may not necessarily have low diffusivities. The blue areas near the left and

430

right boundaries in Figure 8 (a) and (b) are due to the fact that the diffusion flux is too small to be

431

accurately captured within these regions, which is a result of the experimental setup where diffusion is

432

mostly in the horizontal direction pointing to the zero CH2I2 concentration perimeter boundary. Apart

433

from this, two near horizontal blue bands are observed in the R maps that are presumably influenced

434

by fractures. The measured R values of a horizontal fracture may be low due to the reason that in such

435

fractures, where the principal diffusion is in the horizontal direction, vertical diffusion is too small to be

436

accurately captured similar to the effect along the perimeter boundary. However, low R values may

437

also be the result of zero flux symmetry boundary conditions along the center plane between two

438

fractures. The domain that locates between two parallel fractures is a source for CH2I2 for the two

439

fractures, so the center plane between the fractures would have zero flux in the vertical direction due to

440

symmetry. This is a similar phenomenon to the lower Q values around the central axis of the Q

441

volume in Figure4 (a). Apart from the influence of fractures, local heterogeneities are also responsible

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 25

442

for the pattern showing in the R maps, although a more specific explanation may be difficult given

443

tomogram resolution of 5 micrometers/voxel. 3D focused ion beam SEM (FIB-SEM) imaging may be

444

added in the future to identify the local heterogeneities that have low effective diffusion coefficient.

1mm 445 446 447 448 449

(a)

(b)

Figure 8. (a) Local effective ` results of a central longitudinal slice of the Permian Basin sub-plug, same as the one in Figure 2

and Figure 3; (b) local effective ` results of another central longitudinal slice of the Permian Basin sub-plug, perpendicular to the one in (a).

450

4. Conclusions

451

We study the liquid-liquid diffusion in shale using 4D micro-CT imaging. The time-sequenced diffusion

452

tomograms enable us to visualize the dynamic process in 4D, which not only provides detailed

453

information of the diffusion but also makes it possible to identify local heterogeneities of the shale plug.

454

The 4D results revealed that diffusion occurring in the Permian Basin sub-plug is uneven and influenced

455

by matrix heterogeneities that are mainly due to compositional changes within the matrix. Local

456

diffusivity is generally positively related to porosity, i.e. materials with higher porosity have higher

457

effective diffusivity. By measuring the local effective diffusion coefficients, the low porosity materials

458

that cannot be distinguished from density/attenuation in the dry and clean tomograms are clearly

459

identified. Identification of those heterogeneities not only helps us to better characterize the diffusion

460

properties of shale rock but also helps us to evaluate the quality of a shale sample in terms of reservoir

461

rock. Heterogeneity is evident in horizontal layers that are parallel to the bedding, and the degree of

462

heterogeneity varies between layers in the vertical direction, which may be a reflection of the laminated

463

nature of shale rocks. Diffusion anisotropy of the shale sample is evidenced by 3D diffusivity

464

measurement, which shows lower vertical effective diffusion coefficients than horizontal ones in general.

465

Although shale pore structures are unable to be seen at current tomogram resolution, the measured

466

local effective diffusion coefficient enables us to infer smaller scale characteristics over a voxel volume 20 ACS Paragon Plus Environment

Page 21 of 25 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

467

(around one hundred cubic micrometers). For instance, high diffusivity indicates larger porosity, better

468

pore connectivity, and smaller tortuosity.

469

For future work, SEM images and FIB-SEM volumes on selected locations will be obtained to show the

470

nanometer scale pore structures, which will then be linked to their effective diffusivity measured here.

471

Then we are going to run pore scale simulations of diffusion on shale FIB-SEM images with different

472

porosity and calculate their effective diffusivity. The aim is to get the correlation between effective

473

diffusivity and porosity, which in combination with local porosity distribution from micro-CT can be used

474

to map local effective diffusivity for a whole sample plug. Then plug scale simulations of diffusion can be

475

carried out and effective diffusivity for the whole plug can be calculated. In this way, we can accomplish

476

a complete workflow of characterization and diffusion property upscaling for shale matrix from

477

nanometer to plug scale.

478

Acknowledgement

479

This research/project was undertaken with the assistance of resources and services from the National

480

Computational Infrastructure (NCI), which is supported by the Australian Government. We acknowledge

481

funding from the member companies of the ANU/UNSW Digicore research consortium.

482

References

483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501

1. Middleton, R. S.; Gupta, R.; Hyman, J. D.; Viswanathan, H. S., The shale gas revolution: Barriers, sustainability, and emerging opportunities. Applied Energy 2017, 199, 88-95. 2. Zou, C.; Dong, D.; Wang, S.; Li, J.; Li, X.; Wang, Y.; Li, D.; Cheng, K., Geological characteristics and resource potential of shale gas in China. Petroleum Exploration and Development 2010, 37, (6), 641-653. 3. Rivard, C.; Lavoie, D.; Lefebvre, R.; Séjourné, S.; Lamontagne, C.; Duchesne, M., An overview of Canadian shale gas production and environmental concerns. International Journal of Coal Geology 2014, 126, 64-76. 4. Beckwith, R., Shale Gas: Promising Prospects Worldwide. 2011. 5. Vengosh, A.; Jackson, R. B.; Warner, N.; Darrah, T. H.; Kondash, A., A critical review of the risks to water resources from unconventional shale gas development and hydraulic fracturing in the United States. Environmental science & technology 2014, 48, (15), 8334-8348. 6. Wang, Q.; Chen, X.; Jha, A. N.; Rogers, H., Natural gas from shale formation – The evolution, evidences and challenges of shale gas revolution in United States. Renewable and Sustainable Energy Reviews 2014, 30, 1-28. 7. Leu, L.; Georgiadis, A.; Blunt, M. J.; Busch, A.; Bertier, P.; Schweinar, K.; Liebi, M.; Menzel, A.; Ott, H., Multiscale Description of Shale Pore Systems by Scanning SAXS and WAXS Microscopy. Energy & Fuels 2016, 30, (12), 10282-10297. 8. Javadpour, F.; Fisher, D.; Unsworth, M., Nanoscale gas flow in shale gas sediments. Journal of Canadian Petroleum Technology 2007, 46, (10), 55-61.

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

502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549

9. Sommacal, S.; Fogden, A.; Young, B.; Noel, W.; Arena, A.; Salazar, L.; Gerwig, T.; Cheng, Q.; Kingston, A.; Marchal, D.; Perez Mazas, A. M.; Naides, C. H.; Kohler, G.; Cagnolatti, M., 3D Multiscale Imaging of the Distribution of Pores, Organic Matter and Oil in Place in Vaca Muerta Shale Samples. In Unconventional Resources Technology Conference: 2016. 10. Fogden, A.; Latham, S.; McKay, T.; Marathe, R.; Turner, M.; Kingston, A.; Senden, T., Micro-CT Analysis of Pores and Organics in Unconventionals Using Novel Contrast Strategies. In Unconventional Resources Technology Conference, Society of Petroleum Engineers: Denver, Colorado, USA, 2014. 11. Fogden, A.; Arena, A.; Zhang, C.; Carnerup, A.; Goergen, E.; Olson, T.; Cheng, Q.; Middleton, J.; Kingston, A.; Zhang, Y.; Armstrong, R., Combining High-Resolution with Larger Volume Images for Improved Characterization of Mudstone Reservoirs. In Society of Petrophysicists and Well-Log Analysts: 2016. 12. Ojha, S. P.; Misra, S.; Sinha, A.; Dang, S.; Sondergeld, C.; Rai, C., Estimation of Pore Network Characteristics and Saturation-Dependent Relative Permeability in Organic-Rich Shale Samples Obtained From Bakken, Wolfcamp, and Woodford Shale Formations. In Society of Petrophysicists and Well-Log Analysts: 2017. 13. Dadmohammadi, Y.; Misra, S.; Sondergeld, C.; Rai, C., Petrophysical interpretation of laboratory pressure-step-decay measurements on ultra-tight rock samples. Part 2 – In the presence of gas slippage, transitional flow, and diffusion mechanisms. Journal of Petroleum Science and Engineering 2017, 158, (Supplement C), 554-569. 14. Chen, L.; Zhang, L.; Kang, Q.; Viswanathan, H. S.; Yao, J.; Tao, W., Nanoscale simulation of shale transport properties using the lattice Boltzmann method: permeability and diffusivity. Scientific Reports 2015, 5, 8089. 15. Zhang, Y.; Mostaghimi, P.; Fogden, A.; Middleton, J.; Sheppard, A.; Armstrong, R. T., Local diffusion coefficient measurements in shale using dynamic micro-computed tomography. Fuel 2017, 207, 312-322. 16. Hirschfelder, J. O.; Bird, R. B.; Spotz, E. L., The Transport Properties of Gases and Gaseous Mixtures. II. Chemical Reviews 1949, 44, (1), 205-231. 17. He, W.; Lv, W.; Dickerson, J. H., Gas Diffusion Mechanisms and Models. In Gas Transport in Solid Oxide Fuel Cells, Springer International Publishing: Cham, 2014; pp 9-17. 18. Zhu, Q.; D'Agostino, C.; Ainte, M.; Mantle, M. D.; Gladden, L. F.; Ortona, O.; Paduano, L.; Ciccarelli, D.; Moggridge, G. D., Prediction of mutual diffusion coefficients in binary liquid systems with one self-associating component from viscosity data and intra-diffusion coefficients at infinite dilution. Chemical Engineering Science 2016, 147, 118-127. 19. Zhu, Q.; Moggridge, G. D.; D’Agostino, C., A local composition model for the prediction of mutual diffusion coefficients in binary liquid mixtures from tracer diffusion coefficients. Chemical Engineering Science 2015, 132, 250-258. 20. Saripalli, K. P.; Serne, R. J.; Meyer, P. D.; McGrail, B. P., Prediction of Diffusion Coefficients in Porous Media Using Tortuosity Factors Based on Interfacial Areas. Ground Water 2002, 40, (4), 346-352. 21. Petersen, E. E., Diffusion in a pore of varying cross section. AIChE Journal 1958, 4, (3), 343-345. 22. Van Brakel, J.; Heertjes, P., Analysis of diffusion in macroporous media in terms of a porosity, a tortuosity and a constrictivity factor. International Journal of Heat and Mass Transfer 1974, 17, (9), 1093-1103. 23. Epstein, N., On tortuosity and the tortuosity factor in flow and diffusion through porous media. Chemical Engineering Science 1989, 44, (3), 777-779. 24. Whitaker, S., Diffusion and dispersion in porous media. AIChE Journal 1967, 13, (3), 420-427. 25. Alberto Ochoa-Tapia, J.; Stroeve, P.; Whitaker, S., Diffusive transport in two-phase media: spatially periodic models and maxwell's theory for isotropic and anisotropic systems. Chemical Engineering Science 1994, 49, (5), 709-726. 22 ACS Paragon Plus Environment

Page 22 of 25

Page 23 of 25 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

550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595

Energy & Fuels

26. Valdes-Parada, F. J.; Alvarez-Ramirez, J., On the effective diffusivity under chemical reaction in porous media. Chemical Engineering Science 2010, 65, (13), 4100-4104. 27. Mehmani, A.; Prodanović, M.; Javadpour, F., Multiscale, Multiphysics Network Modeling of Shale Matrix Gas Flows. Transport in Porous Media 2013, 99, (2), 377-390. 28. Huang, X.; Bandilla, K. W.; Celia, M. A., Multi-Physics Pore-Network Modeling of Two-Phase Shale Matrix Flows. Transport in Porous Media 2016, 111, (1), 123-141. 29. Yuan, W.; Pan, Z.; Li, X.; Yang, Y.; Zhao, C.; Connell, L. D.; Li, S.; He, J., Experimental study and modelling of methane adsorption and diffusion in shale. Fuel 2014, 117, Part A, 509-519. 30. Wang, J.; Yang, Z.; Dong, M.; Gong, H.; Sang, Q.; Li, Y., Experimental and Numerical Investigation of Dynamic Gas Adsorption/Desorption–Diffusion Process in Shale. Energy & Fuels 2016, 30, (12), 1008010091. 31. Crank, J., Mathematics of diffusion. Oxford university press: London, 1957. 32. Ruckenstein, E.; Vaidyanathan, A.; Youngquist, G., Sorption by solids with bidisperse pore structures. Chemical Engineering Science 1971, 26, (9), 1305-1318. 33. Ojha, S. P.; Misra, S.; Tinni, A.; Sondergeld, C.; Rai, C., Pore connectivity and pore size distribution estimates for Wolfcamp and Eagle Ford shale samples from oil, gas and condensate windows using adsorption-desorption measurements. Journal of Petroleum Science and Engineering 2017, 158, (Supplement C), 454-468. 34. Clarkson, C. R.; Bustin, R. M., The effect of pore structure and gas pressure upon the transport properties of coal: a laboratory and modeling study. 1. Isotherms and pore volume distributions. Fuel 1999, 78, (11), 1333-1344. 35. Pillalamarry, M.; Harpalani, S.; Liu, S., Gas diffusion behavior of coal and its impact on production from coalbed methane reservoirs. International Journal of Coal Geology 2011, 86, (4), 342348. 36. Jing, Y.; Armstrong, R. T.; Mostaghimi, P., Rough-walled discrete fracture network modelling for coal characterisation. Fuel 2017, 191, 442-453. 37. Jing, Y.; Armstrong, R. T.; Ramandi, H. L.; Mostaghimi, P., Coal cleat reconstruction using microcomputed tomography imaging. Fuel 2016, 181, 286-299. 38. Tidwell, V. C.; Glass, R. J., X ray and visible light transmission for laboratory measurement of two‐dimensional saturation fields in thin‐slab systems. Water Resources Research 1994, 30, (11), 2873-2882. 39. Ramandi, H. L.; Mostaghimi, P.; Armstrong, R. T.; Saadatfar, M.; Pinczewski, W. V., Porosity and permeability characterization of coal: a micro-computed tomography study. International Journal of Coal Geology 2016, 154–155, 57-68. 40. Gerami, A.; Mostaghimi, P.; Armstrong, R. T.; Zamani, A.; Warkiani, M. E., A microfluidic framework for studying relative permeability in coal. International Journal of Coal Geology 2016, 159, 183-193. 41. Armstrong, R. T.; Wildenschild, D.; Bay, B. K., The effect of pore morphology on microbial enhanced oil recovery. Journal of Petroleum Science and Engineering 2015, 130, 16-25. 42. Robins, V.; Wood, P. J.; Sheppard, A. P., Theory and Algorithms for Constructing Discrete Morse Complexes from Grayscale Digital Images. IEEE Transactions on Pattern Analysis and Machine Intelligence 2011, 33, (8), 1646-1658. 43. Liu, M.; Mostaghimi, P., High-resolution pore-scale simulation of dissolution in porous media. Chemical Engineering Science 2017, 161, 360-369. 44. Mostaghimi, P.; Liu, M.; Arns, C. H., Numerical simulation of reactive transport on micro-CT images. Mathematical Geosciences 2016, 48, (8), 963-983.

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

596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642

45. Brown, K.; SchlÜTer, S.; Sheppard, A.; Wildenschild, D., On the challenges of measuring interfacial characteristics of three-phase fluid flow with x-ray microtomography. Journal of Microscopy 2014, 253, (3), 171-182. 46. Culligan, K. A.; Wildenschild, D.; Christensen, B. S. B.; Gray, W. G.; Rivers, M. L., Pore-scale characteristics of multiphase flow in porous media: A comparison of air–water and oil–water experiments. Advances in Water Resources 2006, 29, (2), 227-238. 47. Armstrong, R. T.; Porter, M. L.; Wildenschild, D., Linking pore-scale interfacial curvature to column-scale capillary pressure. Advances in Water Resources 2012, 46, 55-62. 48. Liu, M.; Shabaninejad, M.; Mostaghimi, P., Impact of mineralogical heterogeneity on reactive transport modelling. Computers & Geosciences 2017, 104, 12-19. 49. Bultreys, T.; Boone, M. A.; Boone, M. N.; De Schryver, T.; Masschaele, B.; Van Hoorebeke, L.; Cnudde, V., Fast laboratory-based micro-computed tomography for pore-scale research: Illustrative experiments and perspectives on the future. Advances in Water Resources 2016, 95, 341-351. 50. Wildenschild, D.; Sheppard, A. P., X-ray imaging and analysis techniques for quantifying porescale structure and processes in subsurface porous medium systems. Advances in Water Resources 2013, 51, 217-246. 51. Mostaghimi, P.; Armstrong, R. T.; Gerami, A.; Hu, Y.; Jing, Y.; Kamali, F.; Liu, M.; Liu, Z.; Lu, X.; Ramandi, H. L.; Zamani, A.; Zhang, Y., Cleat-scale characterisation of coal: An overview. Journal of Natural Gas Science and Engineering 2017, 39, 143-160. 52. Bultreys, T.; De Boever, W.; Cnudde, V., Imaging and image-based fluid transport modeling at the pore scale in geological materials: A practical introduction to the current state-of-the-art. EarthScience Reviews 2016, 155, 93-128. 53. Blunt, M. J.; Bijeljic, B.; Dong, H.; Gharbi, O.; Iglauer, S.; Mostaghimi, P.; Paluszny, A.; Pentland, C., Pore-scale imaging and modelling. Advances in Water Resources 2013, 51, 197-216. 54. Guerrero Aconcha, U. E.; Kantzas, A., Diffusion of Hydrocarbon Gases in Heavy Oil and Bitumen. In Latin American and Caribbean Petroleum Engineering Conference, Society of Petroleum Engineers: Cartagena de Indias, Colombia, 2009. 55. Liu, Y.; Teng, Y.; Lu, G.; Jiang, L.; Zhao, J.; Zhang, Y.; Song, Y., Experimental study on CO2 diffusion in bulk n-decane and n-decane saturated porous media using micro-CT. Fluid Phase Equilibria 2016, 417, 212-219. 56. Tidwell, V. C.; Meigs, L. C.; Christian-Frear, T.; Boney, C. M., Effects of spatially heterogeneous porosity on matrix diffusion as investigated by X-ray absorption imaging. Journal of Contaminant Hydrology 2000, 42, (2–4), 285-302. 57. Polak, A.; Grader, A. S.; Wallach, R.; Nativ, R., Tracer diffusion from a horizontal fracture into the surrounding matrix: measurement by computed tomography. Journal of Contaminant Hydrology 2003, 67, (1–4), 95-112. 58. Vega, B.; Dutta, A.; Kovscek, A. R., CT Imaging of Low-Permeability, Dual-Porosity Systems Using High X-ray Contrast Gas. Transport in Porous Media 2014, 101, (1), 81-97. 59. Agbogun, H. M. D.; Al, T. A.; Hussein, E. M. A., Three dimensional imaging of porosity and tracer concentration distributions in a dolostone sample during diffusion experiments using X-ray micro-CT. Journal of Contaminant Hydrology 2013, 145, (Supplement C), 44-53. 60. Zhang, Y.; Mostaghimi, P.; Fogdon, A.; Arena, A.; Sheppard, A.; Middleton, J.; Armstrong, R. T., Determination of Local Diffusion Coefficients and Directional Anisotropy in Shale From Dynamic MicroCT Imaging. In Unconventional Resources Technology Conference: 2017. 61. Sheppard, A.; Latham, S.; Middleton, J.; Kingston, A.; Myers, G.; Varslot, T.; Fogden, A.; Sawkins, T.; Cruikshank, R.; Saadatfar, M.; Francois, N.; Arns, C.; Senden, T., Techniques in helical scanning, dynamic imaging and image segmentation for improved quantitative analysis with X-ray micro-CT.

24 ACS Paragon Plus Environment

Page 24 of 25

Page 25 of 25 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

643 644 645 646 647 648

Energy & Fuels

Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 2014, 324, 49-56. 62. Latham, S.; Varslot, T.; Sheppard, A., Image registration: enhancing and calibrating X-ray microCT imaging. Proc. of the Soc. Core Analysts, Abu Dhabi, UAE 2008. 63. Tiab, D.; Donaldson, E. C., Petrophysics : theory and practice of measuring reservoir rock and fluid transport properties. Gulf Professional Pub. Co.: Houston, Tex., 1996.

649

25 ACS Paragon Plus Environment