LABORATORY ESTIMATION AND MODELING OF APPARENT ...

231
The Pennsylvania State University The Graduate School College of Earth and Mineral Sciences LABORATORY ESTIMATION AND MODELING OF APPARENT PERMEABILITY FOR ULTRA-TIGHT ANTHRACITE AND SHALE MATRIX: A MULTI-MECHANISTIC FLOW APPROACH A Dissertation in Energy and Mineral Engineering by Yi Wang © 2017 Yi Wang Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2017

Transcript of LABORATORY ESTIMATION AND MODELING OF APPARENT ...

Page 1: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

The Pennsylvania State University

The Graduate School

College of Earth and Mineral Sciences

LABORATORY ESTIMATION AND MODELING OF APPARENT

PERMEABILITY FOR ULTRA-TIGHT ANTHRACITE AND SHALE MATRIX:

A MULTI-MECHANISTIC FLOW APPROACH

A Dissertation in

Energy and Mineral Engineering

by

Yi Wang

© 2017 Yi Wang

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

August 2017

Page 2: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

ii

The dissertation of Yi Wang was reviewed by the following:

Shimin Liu

Assistant Professor of Energy and Mineral Engineering

Dissertation Adviser

Chair of Committee

Derek Elsworth

Professor of Energy and Mineral Engineering and Geoscience

Zuleima T. Karpyn

Professor of Petroleum and Natural Gas Engineering

Quentin E. and Louise L. Wood Faculty Fellow in Petroleum and Natural Gas

Engineering

Ming Xiao

Associate Professor of Civil and Environmental Engineering

Luis F. Ayala H.

Professor of Petroleum and Natural Gas Engineering

Associate Department Head for Graduate Education

Page 3: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

iii

ABSTRACT

Gas production from unconventional reservoirs such as gas shale and coalbed

methane (CBM) has become a major source of clean energy in the United States. Reservoir

apparent permeability is a critical and controlling parameter for the predictions of shale gas

and coalbed methane (CBM) productions. Shale matrix and tight anthracite are

characterized by ultra-tight pore structure and low permeability at micro- and nano-scale

with gas molecules stored by adsorption. Gas transport in shale and anthracite matrices no

longer always falls into the continuum flow regime described by Darcy’s law, rather a

considerable portion of transport is sporadic and irregular due to the mean free path of gas

is comparable to the prevailing pore scale. Therefore, gas transport in both anthracite and

shale will be a complicated nonlinear multi-mechanistic process. A multi-mechanistic flow

process is always happening during shale gas and CBM production, including Darcy

viscous flow, slip flow, transition flow and Knudsen diffusion and their proportional

contributions to apparent permeability are constantly changing with continuous reservoir

depletion. The complexity of the gas storage and flow mechanisms in ultra-fine pore

structure is diverse and makes it more difficult to predict the matrix permeability and gas

deliverability.

In this study, a multi-mechanistic apparent-permeability model for unconventional

reservoir rocks (shale and anthracite) was derived under different stress boundary

conditions (constant-stress and uniaxial-strain). The proposed model incorporates the

pressure-dependent weighting coefficients to separate the contributions of Knudsen

diffusion and Darcy flow on matrix permeability. A combination of both permeability

Page 4: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

iv

components was coupled with pressure-dependent weighting coefficients. A stress–strain

relationships for a linear elastic gas-desorbing porous medium under hydrostatic stress

condition was derived from thermal-elastic equations and can be incorporated into the

Darcian flow component, serving for the permeability data under hydrostatic stress. The

modeled results well agree with anthracite and shale sample permeability measured data.

In this study, laboratory measurements of gas apparent permeability were

conducted on coal and shale samples for both helium and CO2 injection/depletion under

different stress conditions. At low pressure under constant stress condition, CO2

permeability enhancement due to sorption-induced matrix shrinkage effect is significant,

which can be either clearly observed from the pulse-decay pressure response curves or the

data reduced by Cui et al.’s method. CO2 apparent permeability can be higher than He at

pressure higher than 1000 psi, which may be resulted from limited shale adsorption

capacity. Helium permeability is more sensitive to the variation of Terzaghi effective stress

than CO2 and it is independent of pore pressure. The true effective stress coefficient can be

found two values at low pressure region (<500 psi) and high pressure region (>500 psi).

The negative value indicates Knudsen diffusion and slip flow effect have more impact on

apparent permeability than Terzaghi stress at low pressure.

Additionally, laboratory measurements of gas sorption, Knudsen diffusion

coefficient and coal deformation were conducted to break down the key effects that

influence gas permeability evolution. Adsorption isotherms of crushed anthracite coal

samples was measured using Gibbs adsorption principle at different gas pressures. The

adsorption isotherm result showed that the adsorption capacity at low pressure changes

with a higher rate and thus brings a significant sorption-induced rock matrix

Page 5: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

v

swelling/shrinkage effect. And the isotherm data are important inputs for the Darcy

permeability models. The latter was coupled in the apparent-permeability model as the

Darcy flow component which involves the sorption-induced strain component. Diffusion

coefficients of the pulverized samples were estimated by using the particle method and was

used to calculate the effective Knudsen permeability. The Knudsen diffusion flow

component in the proposed apparent-permeability model was constructed by transforming

Knudsen mass flux into permeability term and used to match the effective Knudsen

permeability based on diffusion data. Increasing trends for all results were performed

during pressure drop down in the result plots. And the modeling result showed very good

agreements with them, giving a solid proof of the availability of Knudsen diffusion

component as part of the proposed model. The results of a series of experimental

measurements of coal deformation with gas injection and depletion revealed that the coal

sorption induced deformation exhibits anisotropy, with larger deformation in direction

perpendicular to bedding than those parallel to the bedding planes. The deformation of coal

is reversible for helium and methane with injection/depletion, but not for CO2. Based on

the modeling results, it was found that application of isotropic deformation in permeability

model can overestimate the permeability loss compared to anisotropic deformation. This

demonstrates that the anisotropic coal deformation should be considered to predict the

permeability behavior of CBM as well as CO2 sequestration/ECBM projects.

Page 6: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

vi

TABLE OF CONTENTS

LIST OF FIGURES ................................................................................................................. ix

LIST OF TABLES ................................................................................................................... xiv

ACKNOWLEDGMENTS ....................................................................................................... xv

Chapter 1 Introduction ........................................................................................................... 1

Chapter 2 Laboratory Investigations of Gas Flow Behaviors in Tight Anthracite and

Evaluation of Different Pulse-Decay Methods on Permeability Estimation .................... 5

1. Introduction ................................................................................................................. 6 2. Background and literature review ............................................................................... 8

2.1 Anthracite-CBM studies ................................................................................... 8 2.2 Compressive storage and sorption effect on coal permeability ........................ 8 2.3 Pulse-decay method for stressed rock permeability estimation ........................ 9 2.4 Slip effect ......................................................................................................... 11

3. Experimental work ...................................................................................................... 13 3.1 Sample Procurement and Preparation .............................................................. 13 3.2 Experimental boundary conditions ................................................................... 14 3.3 Experimental setup and procedure ................................................................... 14 3.4 Helium depletion under constant stress condition ............................................ 15 3.5 Helium depletion in uniaxial strain/in situ condition ....................................... 16 3.6 CO2 depletion permeability measurements ...................................................... 17 3.7 Pulse-decay permeability estimation ................................................................ 17

4. Results and Discussion ............................................................................................... 20 4.1 Pulse-decay pressure curve comparison between helium and CO2 .................. 20 4.2 Helium permeability results under constant and uniaxial strain condition ....... 23 4.3 CO2 gas permeability results with stress-controlled condition ......................... 25 4.4 Intrinsic permeability prediction ...................................................................... 28

5. Conclusion .................................................................................................................. 28 References ........................................................................................................................ 30 Appendix .......................................................................................................................... 40

Chapter 3 Estimation of Pressure-dependent Diffusive Permeability of Coal Using

Methane Diffusion Coefficient: Laboratory Measurements and Modeling ..................... 50

1. Introduction ................................................................................................................. 51 2. Background and literature review ............................................................................... 52

2.1 Gas diffusion in coal ......................................................................................... 52 2.2 Diffusion coefficient measurement techniques ................................................ 54 2.3 Gas diffusive permeability modeling ............................................................... 55

3. Experimental work ....................................................................................................... 57 3.1 Sample procurement and preparation ............................................................... 57 3.2 Experimental setup and procedure ................................................................... 58 3.3 Ad-desorption isotherm measurement and estimation ..................................... 59 3.4 Diffusion coefficient estimation ....................................................................... 60

Page 7: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

vii

4. Results and Discussion ............................................................................................... 61 4.1 Adsorption isotherm results .............................................................................. 61 4.2 Methane diffusion coefficient results ............................................................... 62 4.3 Diffusive permeability results .......................................................................... 65 4.4 Coal type and diffusion .................................................................................... 66

5. Conclusion .................................................................................................................. 67 References ........................................................................................................................ 69 Appendix .......................................................................................................................... 78

Chapter 4 Anisotropy Characteristics of Coal Shrinkage/Swelling and Its Impact on Coal

Permeability Evolution with CO2 Injection ..................................................................... 87

1. Introduction ................................................................................................................. 88 1.1 Deformation of coal matrix and fracture to pressure/stress .............................. 89 1.2 Coal cleat network geometry ............................................................................ 90 1.3 Anisotropy of matrix deformation .................................................................... 91 1.4 Gas mixture sorption ........................................................................................ 92

2. Laboratory experiments .............................................................................................. 93 2.1 Sample preparation ........................................................................................... 93 2.2 Experimental setup ........................................................................................... 94 2.3 Experimental procedure.................................................................................... 94

3. Experimental results and discussion ........................................................................... 96 3.1 Helium injection results .................................................................................... 96 3.2 CO2 injection results ......................................................................................... 97 3.3 Methane injection results .................................................................................. 98 3.4 Comparison of methane and CO2 ..................................................................... 99 3.5 CO2 displacement results .................................................................................. 100

4. Influence of anisotropic deformation on permeability during CO2 injection ............. 102 5. Conclusions and Summary ......................................................................................... 106 References ........................................................................................................................ 107 Appendix .......................................................................................................................... 111

Chapter 5 Apparent Permeability Characterization and Multiple Flow Regimes Modeling

with Helium and CO2 Injection under Hydrostatic Condition ......................................... 120

1. Introduction ................................................................................................................. 121 2. Background and Theories ........................................................................................... 122

1.1 Shale gas flow regimes ..................................................................................... 122 1.2 Permeability measurement approaches ............................................................ 129

3. Experimental work ...................................................................................................... 130 3.1 Sample Procurement and Preparation .............................................................. 130 3.2 Experimental stress boundary conditions ......................................................... 131 3.3 Experimental setup and procedure ................................................................... 131

4. Results and discussion ................................................................................................ 133 4.1 Permeability evolution with flow regimes ....................................................... 134 4.2 Permeability evolution with sorption ............................................................... 135 4.3 Permeability evolution with effective stress ..................................................... 136 4.4 Modeling results of Knudsen diffusion flow .................................................... 139

5. Conclusions ................................................................................................................. 141

Page 8: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

viii

References ........................................................................................................................ 143 Appendix .......................................................................................................................... 153

Chapter 6 Modeling of Apparent Permeability for Ultra-Tight Anthracite and Shale Matrix:

A Multi-mechanistic Flow Approach ............................................................................... 164

1. Introduction ................................................................................................................. 165 2. Background and Theories ........................................................................................... 167

2.1 Diffusion and Molecular flow model ............................................................... 167 2.2 Darcy/viscous flow model ................................................................................ 168

3. Proposed Model .......................................................................................................... 169 3.1 Apparent permeability model with multiple flows ........................................... 169 3.2 Apparent permeability model under in situ condition ...................................... 174 3.4 Permeability model under hydrostatic stress condition .................................... 178

4. Data validation and discussion of proposed apparent permeability model ................. 181 5. Conclusions ................................................................................................................. 184 References ........................................................................................................................ 185 Appendix .......................................................................................................................... 189

Chapter 7 Conclusions ............................................................................................................. 200

Chapter 8 Future Work ............................................................................................................ 204

References ................................................................................................................................ 206

Page 9: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

ix

LIST OF FIGURES

Figure 1 Annual CBM production in USA from1989 to 2013 (EIA, 2014) ............................ 40

Figure 2 Photograph of cylindrical Hazelton anthracite core samples. ................................... 40

Figure 3 Picture of the pulse-decay experimental setup for permeability evolution test. V1

is the valve controlling inlet gas flow from upstream, V2 is the valve controlling outlet

gas flow to downstream and V3/V4 is the valve controlling the confining/axial stress

applied on the sample. ...................................................................................................... 41

Figure 4 Rubber jacket deformation evolution with confining stress (Linear Relationship).

.......................................................................................................................................... 41

Figure 5 Comparison of pulse-decay pressure responses for helium and CO2 injections. The

pressure equilibrium time for helium is at least 30 hours less than that of CO2 at each

pressure step. At high pressure helium pressure can get equilibrium in a relatively

short time while the CO2 pressure yet have a longer. ....................................................... 42

Figure 6 Carbon dioxide pulse-decay pressure curves at three different pore pressure under

constant stress condition in gas depletion process. .......................................................... 43

Figure 7 Helium permeability profiles under constant stress condition, obtained by Brace’s

and Dicker & Smits’s methods respectively. ................................................................... 43

Figure 8 Helium permeability profiles under uniaxial strain condition, obtained by Brace’s

and Dicker & Smits’s methods respectively. ................................................................... 44

Figure 9 Permeability increments of Dicker & Smits's method comparing to Brace's method

at constant pressure at 350 psi. ......................................................................................... 45

Figure 10 Horizontal stress variation for helium depletion under uniaxial strain condition. ... 45

Figure 11 Change in effective horizontal stress with helium depletion under two different

stress-strain conditions: uniaxial strain and constant stress. ............................................ 46

Figure 12 Permeability evolution with helium depletion under both constant stress and

uniaxial strain condition using Dicker & Smits’s method. .............................................. 46

Figure 13 Comparison between three different pulse-decay methods with effective stress. ... 47

Figure 14 Experimental results comparison with anthracite CO2 permeability data by Wang

et al. 2011. ........................................................................................................................ 47

Figure 15 Comparison between intrinsic permeability and apparent permeability obtained

by Cui et al.’s methods. .................................................................................................... 48

Page 10: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

x

Figure 16 Typical CBM production curve with relative volumes of methane and water

through production time. .................................................................................................. 78

Figure 17 Gas desorption, diffusion and Darcy viscous flow in coal matrix. .......................... 79

Figure 18 Picture of the experimental setup for adsorption and diffusion measurement.

Pulverized coal samples are stored in the sample cell. .................................................... 79

Figure 19 A schematic view of sorption/diffusion experimental setup .................................... 80

Figure 20 CH4 absolute adsorption isotherms and Langmuir modeling result of San Juan

sub-bituminous and Pittsburgh #8 bituminous pulverized dry samples. .......................... 80

Figure 21 San Juan coal sample CH4 diffusion coefficient variation with pressure................ 81

Figure 22 Pittsburgh #8 coal sample CH4 diffusion coefficient variation with pressure. ....... 81

Figure 23 Pressure drop curve during diffusion coefficient measurement with initial

pressure at 4.19 MPa. The 10, 20 and 30 minutes recording point are shown as black

“X”. .................................................................................................................................. 82

Figure 24 Comparison of gas molecule behavior in low and high pressure diffusion. ............ 82

Figure 25 San Juan coal sample CH4 permeability evolution correlated to diffusion

coefficient. ........................................................................................................................ 83

Figure 26 Pittsburgh #8 coal sample CH4 permeability evolution correlated to diffusion

coefficient. ........................................................................................................................ 84

Figure 27 Diffusion coefficient data comparison between San Juan and Pittsburgh #8 coal

samples in average values. ............................................................................................... 85

Figure 28 Comparison of CH4 permeability evolution correlated to diffusion coefficient

between San Juan and Pittsburgh #8 coal sample. ........................................................... 86

Figure 29 Conceptual coal cleat system: (a) cubic geometry; (b) matchstick geometry (after

Pan and Connell, 2012) .................................................................................................... 111

Figure 30 Schematic of the experimental setup to measure the sorption induced strain with

gas mixture. ...................................................................................................................... 112

Figure 31 Helium injection and depletion induced strains. ...................................................... 112

Figure 32 Volumetric strain with helium injection for three coal samples. ............................. 113

Figure 33 CO2 injection and depletion induced strains. ........................................................... 114

Figure 34 Methane injection and depletion induced strains..................................................... 115

Figure 35 Volumetric strain comparison of CO2 and methane injection and depletion .......... 115

Page 11: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

xi

Figure 36 Experimental strain results along with Langmuir-type model................................. 116

Figure 37 Volumetric strain by methane displacement by CO2. .............................................. 116

Figure 38 Volumetric strain by methane displacement by CO2 with extended Langmuir

modeled results. ............................................................................................................... 117

Figure 39 Pure sorption induced volumetric strain by CO2 along with Langmuir type model.

.......................................................................................................................................... 117

Figure 40 Conceptual deformation for a single matchstick under constant volume

assumption. ...................................................................................................................... 118

Figure 41 Pure sorption induced linear strains by CO2 along with Langmuir type model. ..... 118

Figure 42 Comparison of permeability behavior for isotropic and anisotropic deformation

conditions. ........................................................................................................................ 119

Figure 43 Average monthly gas production of 15 production wells in Eagleford Shale

Group with flow regimes breakdown from large-scale Darcy/continuum flow to

Knudsen diffusion/molecular flow. .................................................................................. 153

Figure 44 Gas desorption process in nano-scale. During gas production, gas molecules

firstly desorb from the kerogen surface to the pore, and then move outward due to

pressure/concentration drop. Then kerogen tend to shrink, together with the matrix. ..... 154

Figure 45 𝐾𝑛 changing with pressure and average pore size for CO2 at 296 K. Mostly the

flow is lying in the transition flow and slip flow region. ................................................. 154

Figure 46 Photograph of a slice of Marcellus shale drilled core sample cut into a disk with

5.39 mm in thickness and 1 inch in diameter. .................................................................. 155

Figure 47 Schematic view of the pulse-decay experimental setup for permeability evolution

test. V1 is the valve controlling inlet gas flow from upstream, V2 is the valve

controlling outlet gas flow to downstream and V3/V4 is the valve controlling the

confining/axial stress applied on the sample. ................................................................... 155

Figure 48 Pulse-decay experimental data with He and CO2 injection on shale thin disks

(5.39 mm & 3.91 mm) under 1600 psi confining (hydrostatic) stress, processed by

using Cui et al.’s method. ................................................................................................. 156

Figure 49 Permeability data with helium and CO2 injection on shale thin disks (5.39 mm

& 3.91 mm) under 3000 psi confining (hydrostatic) stress, processed by using Cui et

al.’s method. ..................................................................................................................... 157

Figure 50 Comparison between averaged helium and CO2 permeability data of 5.39 mm

sample. ............................................................................................................................. 158

Page 12: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

xii

Figure 51 Helium and CO2 permeability data of 5.39 mm shale disc linearized by plotting

logarithm of permeability versus confining stresses at constant pore pressures to

determine the sensitivity of confining stress to permeability. .......................................... 158

Figure 52 Helium and CO2 permeability data of 5.39 mm shale disc linearized by plotting

logarithm of permeability versus p at constant confining stresses to determine the

sensitivity of pore pressure to permeability. The slopes are positive when pressures

are above ~500 psi and negative when below ~500 psi. .................................................. 159

Figure 53 Permeability versus modified effective stress by fitting experimental data to the

effective stress law for 5.39 mm sample. There are two effective stress coefficients

per sample per gas because of the separated slopes for the permeability-pore pressure

relationships. .................................................................................................................... 160

Figure 54 Permeability values of Knudsen flow permeability model within the same range

of average pore size (5~500 nm) for helium and CO2. .................................................... 161

Figure 55 Shale helium injection data and Knudsen diffusion permeability model

comparison. ...................................................................................................................... 161

Figure 56 Shale CO2 injection data and Knudsen diffusion permeability model comparison.

.......................................................................................................................................... 162

Figure 57 Contribution of Knudsen diffusion flow relative to Darcy viscous flow in terms

of weighting coefficient, as a function of pore pressure and pore radius. ........................ 189

Figure 58 Shale He data and combined flow permeability model comparison. ...................... 190

Figure 59 Shale CO2 data and combined flow permeability model comparison. .................... 191

Figure 60 Combined flow model matching experimental data by varying pore size

parameter in apparent permeability model. ...................................................................... 192

Figure 61 Pore size distribution applied in combined apparent permeability model for

fitting experimental data (screening process)................................................................... 193

Figure 62 The proposed apparent permeability modeling results (He) with lab data

validation for natural-fractured anthracite core sample under uniaxial strain condition

and hydrostatic condition. ................................................................................................ 194

Figure 63 The proposed apparent permeability modeling results (CO2) with lab data

validation for natural-fractured anthracite core sample under uniaxial strain condition

and hydrostatic condition. ................................................................................................ 195

Figure 64 The proposed apparent permeability modeling results (He) with lab data

validation for shale disc samples under uniaxial strain condition and hydrostatic

condition........................................................................................................................... 196

Page 13: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

xiii

Figure 65 The proposed apparent permeability modeling results (CO2) with lab data

validation for shale disc samples under uniaxial strain condition and hydrostatic

condition........................................................................................................................... 197

Page 14: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

xiv

LIST OF TABLES

Table 1 Comparison and evaluation of different pulse-decay methods ................................... 48

Table 2 Proximate analysis of Pennsylvania anthracite coal sample ....................................... 48

Table 3 Experimental data of helium permeability (mD) under stress-controlled and

uniaxial-strain conditions ................................................................................................. 49

Table 4 Experimental data of CO2 permeability (mD) by using different pulse-decay

approaches ........................................................................................................................ 49

Table 5 Proximate analysis of San Juan sub-bituminous and Pittsburgh #8 bituminous coal

samples. ............................................................................................................................ 86

Table 6 Volumetric strain Langmuir-type fit parameters for methane and CO2. ..................... 119

Table 7 Input parameters for Ma et al. permeability model. .................................................... 119

Table 8 Apparent permeability measurement record of Marcellus shale thin disc samples

using pulse-decay method. ............................................................................................... 163

Table 9 Comparison of gas apparent permeability models ...................................................... 198

Page 15: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

xv

ACKNOWLEDGMENTS

First and foremost I want to thank my Ph.D. advisor Dr. Shimin Liu. His progressive mind on

academics and positive attitude on life have been making me moved time after time since 2013

when I started working with him. It has been an honor to be his first student ever. I have been

receiving countless advices and assistance from him, which guided me through the whole Ph.D.

life. He performed to me an excellent example of a great academic advisor.

I’m especially grateful for my Ph.D. dissertation committee members: Dr. Derek Elsworth, Dr.

Zuleima Karpyn and Dr. Ming Xiao. Dr. Elsworth has been enlightening me with solid ideas and

thoughts on my research. Also, he generously provided me the laboratory equipment for rock

permeability tests for my data validation work. Dr. Karpyn has consolidated my knowledge on

porous media flow at my first course taken in Penn State and I received valuable suggestions on

flow behaviors from her. I thank Dr. Xiao for his deliberated effort and patience on my dissertation

modification and timely suggestions for my future career.

I would also like to thank Mr. Rui Zhang and Mr. Long Fan, for creative opinions on research

development and assistance on laboratory works.

Some of the experimental data used in this research were brought from the final technical report

of Illinois Clean Coal Institute (ICCI) funded project (DEV11-2) “Changes in Coal Properties with

Exposure to CO2”. Their support is gratefully acknowledged.

My special appreciation will be given to my wife, Wenting Yue, who keep giving me strong

support all the way, in life and in spirit. And my deepest appreciation and respect belong to my

parents, Xi’an Wang and Ming Chen, and my parents-in-law, Baoguo Yue and Liping Ren. Thanks

to their strong support and confidence in me, I was able to overcome this long and difficulty period

and embrace this coming wonderful closure.

Page 16: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

1

Chapter 1 Introduction

In the United States, the development of coalbed methane (CBM) was initially

encouraged by federal tax incentive during the early 1980s. Since then CBM was

considered as a valuable clean energy resource, and the most recent annual energy report

by US Energy Information Administration (Markowski et al., 2014) reveals an incredible

increment in coalbed methane production from 1989 to 2008. Although after 2009 the

production rate shows a little decline trend, CBM is still an important natural gas

production contributor. Also, shale gas reservoirs play important roles in natural gas supply

in the United States. The most recent annual energy report by US Energy Information

Administration (US EIA) reveals an incredible increment in shale gas production from

2007 to 2013. The gross production from shale gas wells increased from 5 bcf/d in 2007 to

33 bcf/d in 2013, representing up to 40% of US total natural gas production. Pennsylvania

became the second-largest shale gas producing state in 2013 with almost all the growth

coming from the Marcellus play. Similar to coalbed methane reservoirs, shale reservoirs

also have distinctive features compared with traditional reservoirs. Shale gas reservoirs

form within the organic rich source rock but with much smaller pore size and thus lower

permeability. Furthermore, the low permeability and adsorption on the shale retard the

migration of gas to a more permeable reservoir.

During CBM and shale gas productions, the permeability of coal and shale

dynamically changes as a result of pressure drawdown. When pressure decreases, there will

be an increase of the effective stress, defined as the difference between the external stress

Page 17: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

2

and pore pressure, tending to close the aperture of existing fractures (Cui and Bustin, 2005;

Mazumder and Wolf, 2008; Palmer and Mansoori, 1998; Shi and Durucan, 2004;Wang et

al., 2012; Wang et al., 2011). And the pressure drawdown also results in matrix shrinkage

through a thermodynamic energy balance which tends to open the factures and an

enhancement of permeability (Liu and Harpalani, 2013a, 2013b; Pan and Connell, 2007).

The permeability evolution is, therefore, controlled by two competitive effects, namely,

stress induced permeability reduction and matrix shrinkage induced permeability

enhancement during pressure depletion. Additionally, gas flow in ultra-tight rocks is

expected to be influenced by multi-mechanistic flows such as sorption, diffusion, slippage

and, Darcy flows (Javadpour, 2009). The non-Darcy flows could be significant in matrix

because of the extremely tight matrix structure when the mean gas flow path is comparable

with the pore size. Thus, the estimated permeability by assuming only Darcy’s flow may

not be valid for anthracite and shale with non-ideal gases like N2, methane and CO2

(Gensterblum et al., 2014), and the characterization of non-Darcy components raises its

importance for both laboratory measurements and modeling.

Previous investigations and studies contributed significantly to our understanding of

unconventional gas flow dynamics. However, there is still a knowledge gap between the

overall gas deliverability and a holistic understanding of fluid mechanisms in micro- and

nano-pores in shale matrix. A good linkage between non-Darcy flow, such as Knudsen

diffusion in micro- and nano-scale, and Darcy flow in macro- and fracture-scale is still

missing. Firstly, diffusion coefficient measurements are lacking and none of them has made

comparison to the Knudsen diffusion model to improve the understanding of diffusive

flow’s contribution in permeability. The utilization of pulverized samples to measure

Page 18: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

3

diffusion coefficient, trying to represent gas permeability is still considered a good

approach in current researches, regardless of proper stress-type experimental boundary

condition. Moreover, though existing Darcy permeability and apparent permeability

models can provide important knowledge of gas transport process and prediction of gas

permeability, the contribution of different flow components yet got weighted properly. And

none of the apparent permeability models has been incorporated with proper in situ

reservoir condition.

The main hypothesis of this proposed research can be concluded as: gas permeability

in tight matrix is dynamically influenced by gas pressure through multi-mechanistic flow

mechanisms, mainly including sorption, diffusion, slippage and Darcy flow. To address the

problems stated above and follow our main hypothesis, the primary objective of this

research is to characterize and model the multi-mechanistic flow dynamics in tight coal

and shale matrix at different gas pressures. Non-Darcy flow components including

adsorption isotherm and diffusion coefficient, will be experimentally measured on

anthracite coal and shale particles and incorporated into apparent permeability modeling.

Gas permeability has been measured on anthracite coal and shale samples under various

boundary conditions. And finally a multi-mechanistic flow model, i.e. a new apparent

permeability model combining both Darcy and non-Darcy flow effects, will be developed

and validated.

In order to achieve the objective, this research will travel through the following

specific aspects:

1. Laboratory measurements of adsorption isotherm and diffusion coefficient for coal

and shale samples.

Page 19: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

4

2. Laboratory measurements of gas permeability for coal and shale core samples under

constant-stress condition, hydrostatic condition and uniaxial-strain condition.

3. Laboratory measurement of coal deformation with gas injection and depletion.

4. Construction of non-Darcy component model (Knudsen diffusion at this moment),

and evaluation of this model though measured diffusion coefficient.

5. Development of a multi-mechanistic apparent permeability model, combining both

Darcy permeability and Knudsen permeability components, coupled with pressure-

dependent weighting coefficients, under uniaxial-strain condition.

6. Derivation of proposed model under hydrostatic condition.

7. Data validation and result analysis.

Page 20: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

5

Chapter 2 Laboratory Investigations of Gas Flow Behaviors in Tight

Anthracite and Evaluation of Different Pulse-Decay Methods on

Permeability Estimation

Abstract

Permeability evolution in coal is critical for the prediction of coalbed methane

(CBM) production and CO2-enhanced-CBM. The anthracite, as the highest rank coal, has

ultra-tight structure and the gas flow dynamics is complicated and influenced by multi-

mechanistic flow components. Gas transport in anthracite will be a nonlinear multi-

mechanistic process also including non-Darcy components like gas as-/desorption, gas

slippage and diffusion flow. In this study, a series of laboratory permeability measurements

were conducted on an anthracite sample for helium and CO2 depletions under both constant

stress and uniaxial strain boundary conditions. The different transient pulse-decay methods

were utilized to estimate the permeability and Klinkenberg correction accounting for slip

effect was also used to calculate the intrinsic permeability. The helium permeability results

indicate the overall permeability under uniaxial strain condition is higher than that under

constant stress condition because of larger effective stress reduction during gas depletion.

At low pressure under constant stress condition, CO2 permeability enhancement due to

sorption-induced matrix shrinkage effect is significant, which can be either clearly

observed from the pulse-decay pressure response curves or the data reduced by Cui et al.’s

method. But within the same pressure range, there is almost no difference between Brace

et al.’s and Dicker & Smits’s method. Gas slippage effect is also significant at low pressure

for low permeability coal based on the obtained experimental data.

Page 21: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

6

1. Introduction

In United States, the development of coalbed methane (CBM) was initially encouraged

by federal tax incentive during the early 1980s. Since then CBM was considered as a

valuable clean energy resource, and the most recent annual energy report by US Energy

Information Administration (A. Markowski et al., 2014) reveals an incredible increment in

coalbed methane production from 1989 to 2008. Although after 2009 the production rate

shows a little decline trend, CBM is still an important natural gas production contributor.

In US, Pennsylvania is fourth largest coal producing state in the nation in 2014 and the

only state producing anthracite coal. And anthracite coal has a general higher heating value

than other coal types (Coal Age, 2014). The anthracites were known as ultra-tight and also

the highest rank coal with highest fixed carbon content. Additionally, from environmental

standpoint, CO2 sequestration in anthracite coal seams is also attractive due to the high CO2

holding capacity per unit volume/mass. For both anthracite-CBM and CO2-enhanced

CBM, the permeability of coal is one of the key decision-making parameters and thus a

sound knowledge of the permeability evolution for anthracites will be essential.

During CBM production, the permeability of coal dynamically changes as a result of

pressure drawdown. When pressure decreases, there will be an increase of the effective

stress, defined as the difference between the external stress and pore pressure, tending to

close the aperture of existing fractures (Cui and Bustin, 2005; Mazumder and Wolf, 2008;

Palmer and Mansoori, 1998; Shi and Durucan, 2004;Wang et al., 2012; Wang et al., 2011).

And the pressure drawdown also results in coal matrix shrinkage through a thermodynamic

energy balance which tends to open the factures and an enhancement of permeability (S.

Liu & Harpalani, 2013a, 2013b; Pan & Connell, 2007). The permeability evolution is,

Page 22: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

7

therefore, controlled by two competitive effects, namely, stress induced permeability

reduction and matrix shrinkage induced permeability enhancement during pressure

depletion. What’s more, gas flow in anthracites is expected to be influenced by multi-

mechanistic flow dynamics such as sorption, diffusion, slippage and, Darcy flows

(Javadpour, 2009). The non-Darcy flows could be significant in anthracites because of the

extremely tight matrix structure when the mean gas flow path is comparable with the pore

size. Thus, the estimated permeability by assuming only Darcy’s flow may not be valid for

tight anthracites with non-ideal gases like N2, methane and CO2 (Gensterblum et al., 2014),

and the characterization of non-Darcy components raises its importance for both laboratory

measurements and modeling.

In this paper, the transient method “pulse-decay” technique was used to measure the

low permeability on anthracite sample (Brace, Walsh, & Frangos, 1968). However, this

original pulse-decay method has its limitations when applying to coal or other organic-rich

reservoir rocks. For example, it assumes no compressive storage in the rock sample (Hsieh,

Tracy, Neuzil, Bredehoeft, & Silliman, 1981), pure Darcy’s flow components without

sorption effect (Cui et al., 2009) and no gas slippage effect (Heller, Vermylen, & Zoback,

2014). Thus in this study, both pulse-decay approaches with pore compressive storage

effect developed by (Dicker & Smits, 1988) and with sorption effect developed by (Cui et

al., 2009) will be employed along with the classic pulse-decay and Klinkenberg correction

will be introduced to weight the contribution of slip flow, in order to test how non-Darcy

effect would impact the tight coal permeability. Also, the permeability was measured under

various experimental boundary conditions and the influence of different boundary was

discussed in detail.

Page 23: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

8

2. Background and literature review

2.1 Anthracite-CBM studies

Coal is generally considered as a self-source reservoir rock with high gas storage

capacity due to sorption effect. Anthracite, as the highest rank coal, has higher adsorption

capacity for gas storage than lower rank coals (A. K. Markowski, 2014). However,

anthracite coal has a relatively low porosity due to high thermal maturity. Thus the lessons

learned from fluid dynamics in tight-shale may help us to better understand the

permeability evolution of anthracite coal. The past coal permeability studies on anthracites

showed complex permeability behaviors with combined matrix swelling/shrinking and

effective stresses effects (Izadi et al., 2011; Shugang Wang et al., 2011; Yin, Jiang, Wang,

& Xu, 2013). Also, gas transport in anthracites is a multi-mechanistic process including

sorption, diffusion, slip and advection flows. Therefore, a comprehensive characterization

and evaluation of anthracite coal permeability evolution in laboratory scale is critical to

decipher the complexity of gas and coal interactions during CBM/ECBM production.

2.2 Compressive storage and sorption effect on coal permeability

Compressive storage of the reservoir in pulse-decay permeability measurements is

influenced by instantaneous volumetric flow rate change, pressure drop rate and fluid and

reservoir compressibility (Jones, 1997). The original pulse-decay developed by Brace et al.

(1968) assumed no compressive storage effect in rock sample. Hsieh et al. (1981) then

derived a general solution accounting for the compressive storage effect in pulse-decay,

and Dicker and Smits (1988) presented a new model to apply this effect into pulse-decay

method. The significance of this effect depends on the ratio between the compressive

storage inside the sample and in the up-/downstream reservoirs, which means it needs to

Page 24: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

9

be evaluated case by case. Since both Brace’s method and Dicker & Smits’s method have

been widely applied in sample permeability measurements, the feasibility of each method,

in our case, should be deliberately tested for ultra-tight rocks.

As a primary storage mechanism in CBM reservoirs, adsorption is, especially,

necessary for indirect gas content estimation (Hartman, 2008). Gas sorption capacity is

typically influenced by pressure, temperature, microstructure of the rock, and it is further

found that the absorbed amount of gas is proportional to the organic carbon content of the

rock (Hildenbrand, Krooss, Busch, & Gaschnitz, 2006; Pillalamarry, Harpalani, & Liu,

2011; Walls, Diaz, & Cavanaugh, 2012; Zhang, Ellis, Ruppel, Milliken, & Yang, 2012).

For coals, adsorption has indirect influence on gas transport properties (Cui et al., 2009).

Permeability is a factor measuring the ability of fluid flow through a porous medium

following Darcy’s law (Mckernan, Rutter, Mecklenburgh, & Taylor, 2014). During CBM

production, methane molecules desorb from the internal surfaces of matrix resulting a

matrix shrinkage that opens natural cleats and then increase of permeability (Liu and

Harpalani, 2014a; Mitra et al., 2012). In (S. Liu & Harpalani, 2013a), both mechanical

effect and sorption induced strain during reservoir depletion was combined in a sorption-

induced strain model that can be coupled into existing permeability models (S. Liu &

Harpalani, 2013b). This coupled model was tested to be valid for subbituminous coal.

However, the roles of sorption effect on the high rank anthracite permeability has not been

investigate and quantified.

2.3 Pulse-decay method for stressed rock permeability estimation

Significant experimental work has been tried to measure the permeability and its

evolution in coal and other tight rocks. Brace et al. (Brace et al., 1968) firstly introduced

Page 25: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

10

the pulse-decay technique as a transient method derived from Darcy’s law to simply

measure the permeability by applying a pressure difference between two sides of a core

sample. In Table 1, we can see that this classic technique under different laboratory

conditions to estimate the tight rock permeability has been utilized and highly improved

by numerous researchers (Cui et al., 2009; Dicker and Smits, 1988; Jones, 1997; Kamath

et al., 1992; Luffel et al., 1993; Malkovsky et al., 2009; Wang et al., 2011). Dicker and

Smits (1988) proposed a pulse-decay calculation method with pore volume compressive

storage effect correction. However, they didn’t incorporate any adsorption effect and non-

Darcy flow regimes into the calculation to be suitable for the unconventional gas

permeability measurements. Moreover, laboratory estimation of permeability of

unconventional reservoir rocks with adsorption effect has been reported and it has been

traditionally measured either under hydrostatic conditions (Cui et al., 2009; Soeder, 1988)

or in the absence of applied stress (Cui et al., 2009). In (Cui et al., 2009), an approach was

proposed to explicitly include adsorption during pulse-decay method to measure the rock

sample permeability. A sorption capacity term firstly derived by Dicker and Smits (1988)

was implicitly introduced to correct the compressive storage in pore space at different

pressures. Wang et al. (2011) used the original pulse-decay calculation method to measure

the coal permeability and to quantify the sorption amount and sorption-induced strain under

fixed stressed condition. These laboratory work advanced the understandings of the

unconventional gas permeability measurements, but their laboratory conditions are not

representative of true field conditions and consequently, the findings may be subject to

faulty permeability measurements of sorptive-elastic media (Liu and Harpalani, 2014a,

2014b; Mitra et al., 2012).

Page 26: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

11

Mitra et al. (2012) presented a step-wise laboratory permeability experiment under

uniaxial condition, which replicates in situ condition of reservoir by fixing the lateral

dimension and vertical stress. The application of uniaxial strain condition can interpret the

dynamic changes of the state of stress during reservoir depletion (S. Liu & Harpalani,

2014e; J.-Q. Shi, Pan, & Durucan, 2014; Ji Quan Shi & Durucan, 2014). The uniaxial strain

condition is widely accepted as in situ condition for subsurface reservoir development, in

which the lateral boundaries of a reservoir are fixed and do not move, as well as the constant

vertical stress due to the unchanged overburden (Geertsma, 1966; Lorenz, Teufel, &

Warpinski, 1991). A reduction in reservoir pressure, in turn, results in a reduction in stress

acting within and surrounding the reservoir. The horizontal stress acting in a reservoir at

depth is observed to decrease significantly with decreasing reservoir pore pressure (S. Liu

& Harpalani, 2014e). This stress decrease is known from simple theoretical calculations

and has been observed in field for many conventional reservoir formations (Breckels &

Eekelen, 1982; Teufel, Rhett, & Farrell, 1991). In this study, permeability measurements

were conducted on tight anthracite coal samples and different pulse-decay approaches were

applied to figure out the feasibility of each method on unconventional reservoir rocks, with

the evaluation of the permeability data under both constant stress condition and uniaxial

strain condition.

2.4 Slip effect

Note that unconventional reservoir rock has very tight structure, gas flow in matrix is

controlled by multiple flow dynamics including Darcy’s flow, diffusion and gas slippage

(Cui et al., 2009; Javadpour, 2009). Since pulse-decay assumes Darcy flow as the only flow

regime during permeability test, it is critical to at least address the gas slip effect as a

Page 27: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

12

correction to differentiate gas permeability from that of liquid. Klinkenberg (Klinkenberg,

1941) initially identified gas slip effect in porous media flow and introduced apparent

permeability as the corrected gas permeability. At a molecular level, gas molecules collide

with pore walls and tend to slide at the walls instead of losing velocity during gas flow

(Swami, 2012). So it is believed that gas slippage can be significant when the pore throat

size is comparable to the mean free path of gas molecules at given pressure and temperature

(Amyx, Bass, & Whiting, 1960). The equation to predict apparent permeability component

for Klinkenberg effect is described as:

𝑘𝑎 = 𝑘∞ (1 +𝑏𝑘

𝑝) (1)

where 𝑘𝑎 is the corrected permeability, 𝑘∞ is the intrinsic/Darcy permeability, 𝑝 is

pore pressure at each step of experiments and 𝑏𝑘 is the Klinkenberg factor shown as

(Ertekin, King, & Schwerer, 1986; Randolph, Soeder.D.J., & Chowdiah, 1984)

𝑏𝑘 =16𝑐𝜇

𝑤√

2𝑅𝑇

𝜋𝑀 (2)

where c is a constant typically taken as 0.9, 𝜇 is the gas viscosity, M is the fluid

molecular weight, w is the width of pore throat, R is the universal gas constant, and T is

temperature. And when sorptive gas is used in the permeability measurement, its

Klinkenberg factor cannot be directly measured since there is a combination of both

slippage and sorption-induced swelling/shrinking effects. Therefore, the value of b should

be obtained firstly using helium in order to separate slippage effect and shrinkage effect

(Harpalani & Chen, 1997). And the equation used to obtain the slip factor for CO2 is shown

as:

Page 28: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

13

𝑏𝐶𝑂2 =𝜇𝐶𝑂2

𝜇𝐻𝑒√

𝑀𝐻𝑒

𝑀𝐶𝑂2𝑏𝐻𝑒 (3)

where 𝜇𝐶𝑂2 and 𝜇𝐻𝑒 are the kinetic viscosity for CO2 and helium, 𝑀𝐶𝑂2 and 𝑀𝐻𝑒

are the molecular weights for CO2 and helium, and 𝑏𝐶𝑂2 and 𝑏𝐻𝑒 are the Klinkenberg

factors for CO2 and helium, respectively.

Since gas slip flow is happening in tight structure during the measurements, by

obtaining apparent permeability data through the pulse-decay method with sorption,

Klinkenberg correction is able to back estimate the coal intrinsic permeability (J. Li, Liu,

Yao, Cai, & Chen, 2013) which can be further incorporated into existing stress/strain-based

coal permeability models to analysis the effective stress influence on permeability and coal

structural changes and extrapolate the uniaxial strain condition in the laboratory scale.

3. Experimental work

The pulse-decay technique was employed to estimate the permeability of anthracite

coal. The advantages of this method is that the permeability can be calculated directly from

the linear portion of the solution (Kamath et al., 1992) and it is the only option for very low

permeability rocks since it is impossible for maintaining steady-state flow in ultra-low

permeability rocks. With the considerations of gas storage/compressibility and non-Darcy

components, this method will be suitable to estimate tight reservoir rock permeability.

3.1 Sample Procurement and Preparation

Blocks of anthracite coal were obtained from Jeddo coal mine located in Hazleton in

Luzerne County, Pennsylvania. The proximate analysis is summarized in Table 2, and

yielded a fixed carbon percentage of 78.35%. Cylindrical cores were drilled from the

anthracite blocks with one-inch in diameter. Following this, the top and bottom of the

Page 29: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

14

drilled core was trimmed to ~ 2 inches in length and the surfaces were polished to enable

proper placement in the triaxial cell. Two well-prepared samples were shown in Figure 2.

After the sample cores were dried, they were then preserved in a dry and clean plastic

sample bag in a lab-use alloy box for 3 hours before put into triaxial cell, in order to

maintain the samples’ integrity.

3.2 Experimental boundary conditions

To estimate the permeability change under various stress-strain conditions, two

boundary conditions were mimicked in our laboratory, that is, constant stress boundary and

uniaxial strain boundary conditions. In triaxial cell test, the constant stress condition refers

to both the axial and confining stresses were maintained at a constant value throughout the

course of experimental duration. The stresses are generated and maintained by computer-

controlled syringe pumps. The constant stress boundary condition was relatively easy to

achieve. To better replicate in situ condition, the uniaxial strain condition was also

implemented in our measurements. Under this condition, the horizontal physical boundary

of the sample and the vertical stress were maintained constants (Palmer and Mansoori,

1998; Shi and Durucan, 2005). Consequently, the horizontal stress was adjusted to maintain

the zero net horizontal strain with gas injection or depletion. By comparing the results

obtained from both boundary conditions, a quantitative analysis was carried out in this

study.

3.3 Experimental setup and procedure

Figure 3 shows the whole experimental system in our laboratory. The setup includes

a Temco triaxial cell core holder, two ISCO syringe pumps, flowrate and pressure

monitoring and recording systems. The syringe pumps are capable of maintaining or

Page 30: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

15

changing stresses in a controlled manner to the desired stress values. Software panel,

desktop LabVIEW software control panel, is programmed to accurately control the pumps

and record the stresses and injection/ejection volumes of fluid. A rubber jacket is used to

isolate the sample from the confining fluid. The experimental temperature was kept

constant at 296 K (23°C). The sample was sandwiched between two steel loading platens

and then placed inside the core holder. Two comparable fixed volumes, large diameter

Swagelok tubing, serve as the upstream and downstream gas reservoirs. The volumes of

up-/downstream reservoir are 30777 mm3 and 18526 mm3, while the sample volume is

26602 mm3. Two high-accuracy USB-based Omega pressure transducers were installed to

continuously monitor and record pressure-time responses with high sampling rate during

the experimental measurements. The pressure-decay pressure curves will be used to

estimate the permeability. We conducted the pulse-decay with both helium and CO2 as the

test fluids.

3.4 Helium depletion under constant stress condition

As a non-sorbing gas, helium was firstly chosen to test the coal flow property. The

sample was gradually stressed to 1000 psi for the confining stress and 2000 psi for the axial

stress. After the mechanical equilibrium was achieved, the entire system was vacuumed by

a vacuum pump to remove the residual air in the gas flow system. In order to mimic the in

situ gas production procedure, gas depletion was employed. Helium was injected at 950

psi for both upstream and downstream. After the equilibrium was reached, the downstream

was reduced to 800 psi, and in this way a pressure difference (“pressure-pulse”) between

up- and down-stream was created. Then the valves between sample and downstream were

opened to discharge the gas through the sample to downstream reservoir. The permeability

Page 31: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

16

was estimated based on the pressure-time responses. Following this, the downstream

pressure was decreased for next pressure step. Step-wise depletions were carried out with

similar interval for designed number of times up to the final equilibrium pressure ~ 100 psi.

3.5 Helium depletion in uniaxial strain/in situ condition

In order to implement the uniaxial strain/in situ condition, the confining strain was

adjusted to maintain the constant sample horizontal dimension at each pressure step; the

vertical/axial stress were kept constant at 2000 psi. In our syringe pump system, water

inside the pump volume is injected into the triaxial cell, gets compressed and creates

horizontal stress around the sample core by either controlling fluid pressure or fluid volume.

The core sample deformation in horizontal direction can be thus measured by monitoring

how much fluid volume changes inside the pump. During the gas depletion process, the

core sample will tend to shrink because of the increasing effective horizontal stress, and

the fluid volume inside the pump will decrease. As a result, to maintain zero horizontal

deformation, we continuously increased the volume of pump fluid with same amount it

reduced at each pressure steps to make it constant. During the experiment, the deformation

of rubber jacket was subtracted from the overall volumetric strain change in the cell. A

compression test was carried out to obtain a quantitative relationship between rubber jacket

deformation and the applied confining stress. A steel rod with the same diameter and length

was placed in the cell and then the rubber jacket deformation was continuously recorded

with the confining stress from 50 psi to 2500 psi. A linear relationship between the

confining stress and rubber jacket thickness deformation is shown in Figure 4. Taking the

rubber deformation into account, we can accurately maintain the constant horizontal

dimension by adjusting the confining stress. The constant axial stress was achieved by

Page 32: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

17

simply setting the constant pumping pressure in axial direction. Similar to constant stress

condition, the sample was depleted from ~950 to ~100 psi with maintaining uniaxial strain

condition. The sample was initially stressed to 1070 in confining stress and 2000 psi in

vertical stress. During depletion, the permeability at each pressure step was estimated and

the corresponding confining stresses were also recorded.

3.6 CO2 depletion permeability measurements

After the helium cycle, CO2 was used for sorbing gas permeability measurements. In

order to make comparison, the external stresses were set as the same as the helium depletion,

namely, 1000 psi for confining and 2000 psi for axial stresses. And similar depletion steps

were applied from ~900 psi to ~100 psi. The permeability was estimated as each pressure

step.

After the constant stress boundary condition, we tried to replicate the uniaxial strain

condition. Unfortunately, the sample fails during the CO2 depletion which may be

attributed to the “coal weakening” and/or high deviatoric stresses (Harpalani & Mitra,

2009). Thus, we did not report the data for the uniaxial strain CO2 depletion.

3.7 Pulse-decay permeability estimation

The transient pulse-decay approach provides an effective way to measure the gas

permeability of tight rocks. To determine the permeability, the pressure transient equation

was introduced by Brace et al. (1968) shown in equations (4) and (5):

𝑃𝑢(𝑡) − 𝑃𝑑(𝑡) = (𝑃𝑢(𝑡0) − 𝑃𝑑(𝑡𝑜))𝑒−𝛼𝑡 (4)

𝛼 = (𝑘𝑎𝐴

𝜇𝑐𝑔𝐿)(

1

𝑉𝑢+

1

𝑉𝑑) (5)

where 𝑃𝑢(𝑡) and 𝑃𝑑(𝑡) is the pressure of upstream and downstream at time t, 𝑃𝑢(𝑡0) and

Page 33: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

18

𝑃𝑑(𝑡𝑜) is the initial pressure of upstream and downstream respectively, 𝛼 is the slope of

the line when plotting the pressure decay 𝑃𝑢(𝑡) − 𝑃𝑑(𝑡) on semi-log paper against time, A

is the cross-sectional area of the sample, L is the length of sample, 𝑐𝑔 is gas compressibility,

and 𝑉𝑢 and 𝑉𝑑 is the upstream and downstream reservoir volumes respectively. After

pressure data are collected from experiments, the only unknown will be sample

permeability k which can be estimated.

This classic method was still one of the most popular methods to estimate the low-

permeability rocks, but it is questionable whether this method can be directly applied to

compressible testing fluids. Brace et al. (1968) assumed Darcy flow only during the

measurement and no compressive storage in the rock sample for the testing fluids. It was a

good assumption if the testing fluid is water or liquids that can be treated as incompressible

fluid in the testing pressure range. However, the gas is known to be highly compressible

that the storage volume should be corrected to get the real gas transport properties. To

compute both permeability and specific storage of the test sample experimentally, Hsieh et

al. (1981) derived more restrictive analytical solutions of the differential equation

describing the decay curves from the permeability measurement with compressive storage

effect. The general solution of the differential equation for dimensionless pressure

difference and dimensionless time was improved and shown as (Dicker & Smits, 1988):

∆𝑝𝐷 = 2 ∑𝑎(𝑏2+𝜃𝑛

2)−(−)𝑏√(𝑎2+𝜃𝑛2)(𝑏2+𝜃𝑛

2)

𝜃𝑛2(𝜃𝑛

2+𝑎+𝑎2+𝑏+𝑏2)+𝑎𝑏(𝑎+𝑏+𝑎𝑏)∞𝑛=1 ×𝑒(−𝜃𝑛

2𝑡𝑑) (6)

where, 𝑎 and 𝑏 is the ratio of sample’s storage capacity to that of upstream reservoir and

downstream reservoir, and 𝜃𝑛 is the nth root of the following equation:

Page 34: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

19

𝑡𝑎𝑛𝜃 =(𝑎+𝑏)𝜃

𝜃2−𝑎𝑏 (7)

where 𝑎 =𝑉𝑝

𝑉𝑢, 𝑏 =

𝑉𝑝

𝑉𝑑 and 𝑉𝑝 is the pore volume of rock sample.

To simplify the above method, Jones (1997) introduced a factor 𝑓 as follows:

𝑓 =𝜃2

𝑎+𝑏 (8)

And the original pulse-decay equation turns into:

𝛼 = (𝑓𝑘𝑎𝐴

𝜇𝑐𝑔𝐿)(

1

𝑉𝑢+

1

𝑉𝑑) (9)

Then the measured sample permeability becomes:

𝑘𝑎 = 𝛼𝜇𝑐𝑔𝐿

𝑓𝐴(1

𝑉𝑢+

1

𝑉𝑑) (10)

Another feature of the organic-bearing rocks is adsorption. The Dicker & Smits’s

method can correct the compressive storage, but it could not handle the loss of adsorbed

gas during the gas injection because the adsorbed gas is no longer in gaseous phase. In

order to extend the pulse-decay technique for sorptive and tight material, Cui et al. (2009)

presented a new approach to estimate permeability with both pore compressive storage

effect and sorption effect for organic-bearing rocks. Corresponding to the effective porosity

of core sample, an effective adsorption porosity term is introduced to account for the

contribution off gas molecule adsorption. Langmuir model was used to quantify the gas

adsorption volume as a function of pressure (Langmuir, 1918) and mathematically

described as follows:

𝑉𝑎 =𝑉𝐿𝑝

𝑝𝐿 + 𝑝 (11)

where, 𝑉𝑎 is the gas adsorbed volume, 𝑉𝐿 is the Langmuir volume and 𝑝𝐿 is the Langmuir

Page 35: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

20

pressure. So the sample storage capacity ratio in the Cui et al.’s approach becomes:

𝑎 =𝑉𝑝 (1 +

𝜙𝑎

𝜙 )

𝑉𝑢 (12)

𝑏 =𝑉𝑝 (1 +

𝜙𝑎

𝜙 )

𝑉𝑑 (13)

where, 𝜙𝑎 = 𝜌𝑠(1−𝜙)𝑝𝐿𝑉𝑎

𝜌𝑉𝑠𝑡𝑑𝐶𝑔(𝑝𝐿+𝑝)2, 𝜙 is the matrix porosity of rock sample, 𝜌 and 𝜌𝑠 is the molar

density of gas and the skeleton density of porous sample respectively, 𝑉𝑠𝑡𝑑 is the molar

volume of gas at standard pressure and temperature (i.e. 273.15 K and 101,325 Pa). In this

study, we compared the estimated permeabilities by different aforementioned approaches

including classic Brace approach, Dicker & Smits method and Cui et al. approach, and

recommendations were made for the tight anthracites.

4. Results and Discussion

4.1 Pulse-decay pressure curve comparison between helium and CO2

Figure 5 shows the helium and CO2 pulse-decay pressure response curves measured at

each gas pressure step during constant stress conditions. Due to the fact that equilibrium

time is extremely long (may last for more than 1 week at 100 psi gas pressure) at low

pressure, we only took partial pressure-decay curve to estimate the permeability. This is an

advantage feature of the transient method. From Figure 5, we found that the time required

to approach the equilibrium decreases as increase of injection pressure for both helium and

CO2. Considering helium as a non-sorbing gas, only effective stress effect influences the

permeability evolution, and the permeability are expected to increase with elevated pore

pressure at which the effective stress reduced correspondingly.

Page 36: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

21

For CO2 gas, beside the effective stress effects, the sorption process, happened on the

internal surface of coal matrix, influences the structure of coal matrix and thus the gas

deliverability. Driven by multiple mechanisms, the permeability of CO2 is expected to

follow different variation trend compared to helium which will be directly captured and

visualized from the pressure curves. From Figure 5, the CO2 gas pressure equilibrium time,

as we can observe, is generally more than that for helium, which physically indicates CO2

transport slower than helium under similar conditions. And the pressure decay rate is lower

between 300~500 psi than other pressures, which indicates a lower permeability in this

pressure range.

If we make a careful comparison with helium and CO2 pressure decay curves, a few

notable findings can be observed qualitatively. According to Figure 5, the overall time for

helium permeability measurement is just as half as the time for CO2, and even that a few

helium curves already get equilibrium while most the CO2 still on its half way at most

pressure steps. For example, when the pseudo equilibrium pressure is about 300 psi

(upstream pressure about 400 plus and downstream pressure about 100 plus for both gases),

the helium pressure difference between upstream and downstream reservoir change from

340 to 98 psi within less than 1 x 105 seconds while it took almost 2 x 105 seconds for the

CO2 pressure difference to only vary from 330 to 196 psi. With this comparison, we can

conclude that it requires much more time for the equilibrium for CO2 than for helium to

pass through the rock sample at same condition, which indicates the helium gas

permeability is higher than the CO2 gas permeability. This also indicates the coal

permeability is a gas type dependent property. For non-sorbing gas, the total free gas keeps

constant in the whole system. However, for anthracite coal and other sorptive reservoir

Page 37: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

22

rocks, the gas ad-/desorption process can either store in or produce gas from the matrix,

dominating the flux in/out the matrix. In Figure 5, we can observe several obvious

asymmetries for CO2 depletion. For example, when the pseudo equilibrium pressure at 550

psi, the upstream pressure drops in a higher speed from 640 to 600 psi while the

downstream pressure hardly has any increment larger than a few psi. This indicates that

the absolute free gas quantify decrease as time elapse. This lost gas quantity should be

considered during the permeability measurement.

In order to qualitatively analyze the sorption effect on the pressure decay cures, we

repeat three pressure decays for the same coal sample under the same experimental

condition where we allowed fully equilibrium as shown in Figure 6. The sample was

depleted from ~700 psi. The dot lines represent the pseudo equilibrium pressure at each

pressure step. One interesting observation is that the equilibrium pressure slightly increases

with time elapse at 410 and 120 psi, which indicates there is extra gas mass influx to the

free gas phase. This may be attributed to the slow gas desorption with gas depletion.

Because the desorption is a slower process than pressure-driven flow. Conceptually, the

equilibrium takes two coherent processes: one is the pressure drive equilibrium where the

pressure equilibrium between up- and down volume happens relatively fast; the other is

desorption-driven flow which happens relative slow. This desorption-driven flow makes

the tails of pressure equilibrium slightly upwards instead of maintaining stationary at low

pressures. And this effect is hardly seen at high pressure because of the nature of the

sorption behavior that sorption reach the plateau at high pressures for coals (Busch,

Gensterblum, Krooss, & Littke, 2004; Harpalani, Prusty, & Dutta, 2006).

Page 38: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

23

If we look into different scale of this phenomenon, we may see that the ad-/de-sorption

effect has the real influence on the permeability evolution during reservoir pressure change.

For unconventional reservoirs, the adsorption effect provides a large gas storage capacity

while the desorption process leads to a significant increment on the total non-Darcy flux,

and thus may control late-time production reservoir behavior, if desorption/diffusion is the

rate-limiting step (Cui et al., 2009; Yi et al., 2008). Besides, the shrinkage of matrix due to

desorption effect also results in the rock matrix deformation and, consequentially,

permeability change (H. Kumar, Elsworth, Liu, Pone, & Mathews, 2012; H. Kumar,

Elsworth, Mathews, Liu, & Pone, 2014; S. Liu & Harpalani, 2013a; Shugang Wang,

Elsworth, & Liu, 2012b).

4.2 Helium permeability results under constant and uniaxial strain condition

As the section 3.4 mentioned, the compressive storage could cause the error in the

permeability estimation. We quantitatively analyzed how the compressive storage

influence the permeability estimation here. Since helium is a non-sorbing gas, helium

permeability was estimated under constant stress condition obtained by Brace’s method

(without compressive storage correction) and Dicker & Smits’s method (with compressive

storage correction). In order to apply Dicker & Smits’s method, the initial porosity is

required and was assumed to be 8% (Gan, Nandi, & Walker, 1972; Rodrigues & Lemos De

Sousa, 2002). For both methods, the permeabilities were estimated and plotted in Figure 7.

The permeability decrease from ~0.01 to ~ 0.0002 millidarcy (md) with pore pressure

decreasing from 900 to 100 psi. Although Dicker & Smits’s method includes the

compressive storage effect, there is almost no difference between the two permeability

profiles. This is not surprising since the anthracite coal has really low porosity. This has

Page 39: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

24

been confirmed for the uniaxial strain condition with the results shown in Figure 8. The

permeabilities estimated by both methods were very similar.

In order to test how the contribution of pore volume compressive storage varies with

pore volume, we calculated the permeability difference in percentage between Brace’s

method and Dicker & Smits’ method, within a porosity range between 5% ~ 30%. Figure

9 shows the permeability difference between Dicker & Smits’s method and original Brace’s

method. The difference between the two methods increases with the increase of sample

porosity. With the porosity below 10%, Dicker & Smits’s estimated values are at most 5%

larger than Brace’s method. When the sample porosity increases to 30%, this difference

can become up to 12%. Note that in this study the sample volume is not so different with

up-/downstream reservoir volume, and the pore volume takes up only 6% of upstream

volume and 11% of downstream volume. Consequently, we may safely assume that at a

relative low pressure for non-sorbing gas, the pore compressive storage effect, in our case

when the sample volume is quite small compared to up-/downstream volume, is

insignificant and thus there is no significant difference between Brace’s method and Dicker

& Smits’s method. Either method can be used predict the permeability at both constant

stress condition and uniaxial strain condition. For other type of ultra-tight rocks, the

compressive storage may not be significant due to the low porosity nature. Therefore, it

would be a good assumption that for shale and tight reservoirs, the Brace’s method can be

safely applied for the non-sorbing gas permeability estimation at relatively low pressures.

Now we compare the permeability results obtained by Brace’s method under two

boundary conditions. Differed from stress-controlled condition, horizontal stress decreased

from 1070 to 770 psi for pore pressure depletion from 850 to 150 psi, as shown in Figure

Page 40: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

25

10. The effective horizontal stress for both boundary conditions were calculated and shown

in Figure 11. With depletion, the increase of effective horizontal stress for uniaxial strain

condition is slower than for constant stress condition. The reason is simply because

horizontal stress under uniaxial strain condition kept decreasing to maintain the zero

horizontal strain resulting in a relatively slow effective horizontal stress increase. This

phenomenon has been observed for low rank coals in our previous studies (Liu and

Harpalani, 2014c; Mitra et al., 2012). As expected, the permeability under uniaxial strain

conditions was found to be higher than the constant stress condition as shown in Figure 12

and Table 3. This results well correlate to the effective stress variations as Figure 11. With

increase of effective stress, the permeability decreases with depletion.

4.3 CO2 gas permeability results with stress-controlled condition

In order to test influence of different pulse-decay calculation methods for sorbing gas,

CO2 depletion was conducted under constant stress condition. Based on the pressure

response curves obtained during the experiment, comparisons were made between the three

pulse-decay calculation methods, original pulse-decay (Brace et al., 1968), modified

method with pore compressive storage effect (Dicker & Smits, 1988) and modified method

with sorption effect (Cui et al., 2009). The results were shown in Figure 13. The

permeabilities were calculated with three methods. In order to take the sorption into

consideration in Cui et al.’ method, the sorption data were measured in our lab, including

Langmuir pressure(𝑝𝐿 = 200 𝑝𝑠𝑖), Langmuir volume (𝑉𝐿 = 24 𝑚𝐿/𝑔), sample length

(𝐿 = 52.5 𝑚𝑚), temperature ( 𝑇 = 295.15 𝐾 ). And gas properties and other pressure-

temperature dependent parameters were calculated by following ASTM standard methods.

Page 41: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

26

In Figure 13, the calculated permeabilities using three different pulse-decay methods

behave somewhat similar. The permeability initially declines sharply with pressure-

depletion from 850 to 400 psi and then start to respond. It is well known that the gas

permeability in coal was simultaneously controlled by the effective stress and

microstructural change due to sorption known as matrix shrinkage (Cui and Bustin, 2005;

Palmer and Mansoori, 1998; Shi and Durucan, 2005). The initial permeability decrease was

attributed to the effective increase as shown in the figure. Although the desorption-induced

shrinkage happens in these high pressures, but the dominate effect is the increase of

effective stress tending to narrow the gas flow channels. When the pore pressure keeps

decreasing, a significant portion of CO2 will desorb from the matrix and this sorption

induced permeability increase will dominate the flow behaviors in the low pressures.

Because of matrix shrinkage effect, the permeability tends to increase though the effective

stress increase. This CO2 permeability behavior in pulse-decay test matches previous

laboratory measurements on tight coal samples done by several researchers (Izadi &

Elsworth, 2013; J. Li et al., 2013; Shugang Wang et al., 2011). However, the permeability

did not recover to or exceed its original values for anthracite, which is different from some

of the low-rank coals (Liu and Harpalani, 2013b; Mitra et al., 2012). This might be due to

the tight structure of anthracite and the sorption induced matrix shrinkage is comparatively

less than bituminous coals.

Comparing to the permeability data from original pulse-decay mothed, the result from

Dicker & Smits’s method still shows almost no difference with Brace’s method, indicating

the pore compressive effect at low pressure is also less significant with CO2 gas, as we can

see from the values in Table 4. On the other hand, at low pressure less than 100 psi, there

Page 42: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

27

is an obvious enhancement in Cui et al.’s model. By considering the sorption effect, the

Cui et al.’s method estimates higher permeability than others two methods, especially at

low pressure. For example, the estimated permeability Cui et al.’s approach is 28% higher

than the other two at 30 psi pore pressure. The reason why the estimated permeability

deviation is elevated is that the sorption effect becomes more significant at low pressure.

And clearly this sorption process during gas depletion have a positive influence on the

permeability. The contribution of sorption to overall multi-mechanistic flow is thus

important and the pulse-decay method with sorption correction significantly benefits

experimental characterization of tight rock permeability.

In order to test the availability of our permeability results, previous experimental data

on Pennsylvania anthracite coal sample was used in this study. Figure 14 shows the

permeability results comparison between the data from Wang et al. (2011) and this study.

In their work, the coal sample was collected from the Northumberland Basin, Mount

Carmel in Pennsylvania. Similar experimental setup was used for pulse-decay and a

constant confining/axial stress of 6 MPa (870) was applied. The method they followed is

Brace et al.’s method. Compared to our data with same method, the permeability value of

theirs is very close to our results, and the permeability trend is also a similar “check-shape”.

Overall their values are higher than ours with limited amount. This may due to the fact they

used lower boundary stresses, since less confining/axial stress generally resulted in less

effective stress and higher permeability in this case. Although other factors such as mineral

contents, adsorption capacity, physical properties etc. may contribute to the difference in

this comparison, the anthracite permeability data in this study was proved to be viable.

Page 43: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

28

4.4 Intrinsic permeability prediction

Generally speaking, rock’s intrinsic permeability obtained by Klinkenberg

correction can be lower than the apparent permeability, because slip effect is such a

phenomenon showing that gas permeability potentially is higher than pure liquid

permeability at the same condition. The influence of Klinkenberg effect on low

permeability reservoirs will increase with the reduction of gas pressure (G. Wang, Ren,

Wang, & Zhou, 2014). To apply the Klinkenberg correction, the average width of pore

throat in anthracite coal is assumed 0.001 μm in this study (Halliburton Company, 2007).

The apparent permeability data obtained by the above experiments were reduced by

Klinkenberg correction to the intrinsic permeability. We took Cui et al.’s method result

only since the other two methods follow the same behavior and the result was shown in

Figure 15. The intrinsic permeability starts to deviate from the measured apparent

permeability from 180 psi. There is a 20% reduction of permeability value comparing to

the apparent permeability at 30 psi. On the contrary, the intrinsic permeability is almost the

same with the apparent permeability above 180 psi. This phenomenon consists with the

most recently findings for the tight shales that the slip flow play increasingly important

role at low pressures and is minimal at high pressures (Civan, Rai, & Sondergeld, 2011;

Javadpour, 2009; Javadpour, Fisher, & Unsworth, 2007). Therefore, based on the obtained

results, we may see the importance of non-Darcy flow on tight rock permeability

measurement and analysis.

5. Conclusion

A series of experimental studies and theoretical analyses on permeability of Hazelton

anthracite core sample under different stressed controlled conditions has been presented.

Page 44: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

29

The results show that the permeability evolution of studied coal is pressure-boundary

dependent and it is simultaneously controlled by the effective stress profile and the sorption

process which tends to alter the microstructure of coal. Three different pulse-decay

calculation methods for conventional and unconventional gas permeability are utilized and

compared. Based on the work completed, the following conclusions are made and

summarized:

1. Under uniaxial strain condition, the applied horizontal stress linearly decreased with

gas pressure depletion, which consists with traditional oil/gas reservoirs. Because of

the horizontal stress loss, the permeability under constant stress condition with helium

depletion was found to be less than the results under uniaxial strain condition.

2. The sorption induced matrix shrinkage plays important role on the permeability

enhancement at low pressure for the anthracite coal. But it is not strong enough to

compensate the stress effect as the bituminous coal did.

3. For the pulse-decay method, the contribution of ad-/desorption can be clearly observed

from the pressure respond curves. And this effect is stronger at low pressure than high

pressures.

4. Comparing to Brace’s method, Dicker & Smits’s method incorporates the pore

compressive storage effect only and it may have more influence on gas permeability at

pressure higher than 1000 psi. On the other hand, these two methods can be identical

and both valid at relatively low pressure. For simplicity, the Brace’s method gives

reliable data at the range of investigated pressures for tight and ultra-tight rocks.

5. Cui et al.’s method gives an obvious enhancement for the contribution of sorption effect

in permeability calculation, providing a good direction of how the gas sorption effect

Page 45: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

30

can help to predict apparent permeability.

6. Based on the data observation, at extremely low permeability the Klinkenberg effect

becomes significant, and gas slippage is considered to be an important effect when

predicting unconventional reservoir gas permeability due to multiple flow mechanism.

The characterization of gas slip flow, along with sorption and other non-Darcy flow

components, plays an important role in both production analysis and laboratory

measurement. It is worthwhile to pay more attention on how to measure and calculate

the apparent permeability in a laboratory scale with so many non-Darcy components in

future study.

References

Aminian, K., & Ameri, S. (2004). Type Curves for Coalbed Methane Production

Prediction. 2004 SPE Eastern Regional Meeting. http://doi.org/10.2523/91482-MS

Amyx, J. W., Bass, D. M. J., & Whiting, R. L. (1960). Petroleum Reservoir Engineering.

An, F. H., Cheng, Y. P., Wu, D. M., & Wang, L. (2013). The effect of small micropores

on methane adsorption of coals from Northern China. Adsorption, 19(1), 83–90.

http://doi.org/10.1007/s10450-012-9421-3

Brace, W. F., Walsh, J. B., & Frangos, W. T. (1968). Permeability of granite under high

pressure. Journal of Geophysical Research, 73(6).

Breckels, I. M., & Eekelen, H. a. M. (1982). Relationship Between Horizontal Stress and

Depth in Sedimentary Basins. Journal of Petroleum Technology, 34(9), 2191–2199.

http://doi.org/10.2118/10336-PA

Busch, A., Gensterblum, Y., Krooss, B. M., & Littke, R. (2004). Methane and carbon

dioxide adsorption–diffusion experiments on coal: upscaling and modeling.

International Journal of Coal Geology, 60(2–4), 151–168.

http://doi.org/10.1016/j.coal.2004.05.002

Cai, Y., Liu, D., Pan, Z., Yao, Y., Li, J., & Qiu, Y. (2013). Pore structure and its impact

on CH4 adsorption capacity and flow capability of bituminous and subbituminous

coals from Northeast China. Fuel, 103, 258–268.

Page 46: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

31

http://doi.org/10.1016/j.fuel.2012.06.055

Choi, J.-G., Do, D. D., & Do, H. D. (2001). Surface Diffusion of Adsorbed Molecules in

Porous Media: Monolayer, Multilayer, and Capillary Condensation Regimes.

Industrial & Engineering Chemistry Research, 40(19), 4005–4031.

http://doi.org/10.1021/ie010195z

Civan, F., Rai, C., & Sondergeld, C. (2011). Shale-gas permeability and diffusivity

inferred by improved formulation of relevant retention and transport mechanisms.

Transport in Porous Media, 86(3), 925–944. http://doi.org/10.1007/s11242-010-

9665-x

Clarkson, C., & Bustin, R. (1999). 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, di, 1333–1344.

Clarkson, C., Pan, Z., Palmer, I., & Harpalani, S. (2010). Predicting sorption-induced

strain and permeability increase with depletion for coalbed-methane reservoirs. SPE

Journal, (March).

Clarkson, C. R., & Bustin, R. M. (1999). The effect of pore structure and gas pressure

upon the transport properties of coal: a laboratory and modeling study. 2. Adsorption

rate modeling. Fuel, 78(11), 1345–1362. http://doi.org/10.1016/S0016-

2361(99)00056-3

Clarkson, C. R., Solano, N., & Bustin, R. (2013). Pore structure characterization of North

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

intrusion. Fuel, 103, 606–616. http://doi.org/10.1016/j.fuel.2012.06.119

Coal Age - DEC 2014. (2014).

Coppens, M. (1999). The effect of fractal surface roughness on diffusion and reaction in

porous catalysts – from fundamentals to practical applications. Catalysis Today,

53(2), 225–243. http://doi.org/10.1016/S0920-5861(99)00118-2

Coppens, M. O., & Dammers, A. J. (2006). Effects of heterogeneity on diffusion in

nanopores - From inorganic materials to protein crystals and ion channels. Fluid

Phase Equilibria, 241(1–2), 308–316. http://doi.org/10.1016/j.fluid.2005.12.039

Crank, J. (1957). The mathematics of diffusion.

Cui, X., Bustin, A., & Bustin, R. (2009). Measurements of gas permeability and

diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 40479. R

Cui, X., Bustin, A. M. M., & Bustin, R. M. (2009). Measurements of gas permeability

and diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 9(3), 208–223. http://doi.org/10.1111/j.1468-8123.2009.00244.x

Cui, X., & Bustin, R. M. (2005). Volumetric strain associated with methane desorption

and its impact on coalbed gas production from deep coal seams. AAPG Bulletin,

89(9), 1181–1202. http://doi.org/10.1306/05110504114

Cui, X., Bustin, R. M., & Dipple, G. (2004). Selective transport of CO2, CH4, and N2 in

Page 47: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

32

coals: insights from modeling of experimental gas adsorption data. Fuel, 83(3), 293–

303. http://doi.org/10.1016/j.fuel.2003.09.001

Darabi, H., Ettehad, a., Javadpour, F., & Sepehrnoori, K. (2012). Gas flow in ultra-tight

shale strata. Journal of Fluid Mechanics, 710, 641–658.

http://doi.org/10.1017/jfm.2012.424

Dicker, A. I., & Smits, R. M. (1988). A practical approach for determining permeability

from laboratory pressure-pulse decay measurements. In SPE International Meeting

on Petroleum Engineering, Tianjin, China,1-4 Nov,1998. Tianjin, China.

Do, D. D. (1998). Adsorption Analysis: Equilibria and Kinetics (Vol. 2). Imperial

College Press. http://doi.org/10.1142/9781860943829

Drazer, G., Chertcoff, R., Bruno, L., & Rosen, M. (1999). Concentration dependence of

diffusion–adsorption rate in activated carbon. Chemical Engineering Science, 54,

4285–4291.

Dutta, A. (2009). Multicomponent Gas Diffusion and Adsorption in Coals for Enhanced

Methane Recovery.

EIA. (2015). Marcellus Region Drilling Productivity Report, (March), 2015.

Ertekin, T., King, G., & Schwerer, F. (1986). Dynamic Gas Slippage: A Unique Dual-

Mechanism Approach to the Flow of Gas in Tight Formations. SPE Formation

Evaluation. http://doi.org/10.2118/12045-PA

Gan, H., Nandi, S. P., & Walker, P. L. (1972). Nature of the porosity in American coals.

Fuel, 51(4), 272–277. http://doi.org/10.1016/0016-2361(72)90003-8

Geertsma, J. (1966). Problems of rock mechanics in petroleum production engineering.

1st ISRM Congress, 3(59), 585–594.

Gensterblum, Y., Ghanizadeh, A., & Krooss, B. M. (2014). Gas permeability

measurements on Australian subbituminous coals: Fluid dynamic and poroelastic

aspects. Journal of Natural Gas Science and Engineering, 19, 202–214.

http://doi.org/10.1016/j.jngse.2014.04.016

Halliburton Company. (2007). Chapter 04 - Reservoir Analysis. Coalbed Methane

Princples and Practices, (June 2007), 191–282.

Harpalani, S., & Chen, G. (1997). Influence of gas production induced volumetric strain

on permeability of coal. Geotechnical and Geological Engineering, 15(4), 303–325.

http://doi.org/10.1007/BF00880711

Harpalani, S., & Mitra, A. (2009). Impact of CO2 Injection on Flow Behavior of Coalbed

Methane Reservoirs. Transport in Porous Media, 82(1), 141–156.

http://doi.org/10.1007/s11242-009-9475-1

Harpalani, S., Prusty, B., & Dutta, P. (2006). Methane/CO2 sorption modeling for

coalbed methane production and CO2 sequestration. Energy & Fuels, 20(4), 1591–

1599.

Hartman, R. (2008). Recent advances in the analytical methods used for shale gas

Page 48: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

33

reservoir gas-in-place assessment. In AAPG Annual Convention, San Antonio,

Texas, 20-23 Apr, 2008 (Vol. 40317).

He, W., Lv, W., & Dickerson, J. (2014). Gas Diffusion Mechanisms and Models. In Gas

Transport in Solid Oxide Fuel Cells (pp. 9–17). http://doi.org/10.1007/978-3-319-

09737-4

Heller, R., Vermylen, J., & Zoback, M. (2014). Experimental investigation of matrix

permeability of gas shales. AAPG Bulletin, 98(5), 975–995.

http://doi.org/10.1306/09231313023

Hildenbrand, a., Krooss, B. M., Busch, a., & Gaschnitz, R. (2006). Evolution of

methane sorption capacity of coal seams as a function of burial history — a case

study from the Campine Basin, NE Belgium. International Journal of Coal Geology,

66(3), 179–203. http://doi.org/10.1016/j.coal.2005.07.006

Hsieh, P. a., Tracy, J. V., Neuzil, C. E., Bredehoeft, J. D., & Silliman, S. E. (1981). A

transient laboratory method for determining the hydraulic properties of “tight”

rocks—I. Theory. International Journal of Rock Mechanics and Mining Sciences &

Geomechanics Abstracts, 18(3), 245–252. http://doi.org/10.1016/0148-

9062(81)90979-7

Hubbert, M. K. (1957). Darcy’S Law and the Field Equations of the Flow of

Underground Fluids. International Association of Scientific Hydrology. Bulletin,

2(1), 23–59. http://doi.org/10.1080/02626665709493062

Izadi, G., & Elsworth, D. (2013). The effects of thermal stress and fluid pressure on

induced seismicity during stimulation to production within fractured reservoirs.

Terra Nova, 25(5), 374–380. http://doi.org/10.1111/ter.12046

Izadi, G., Wang, S., Elsworth, D., Liu, J., Wu, Y., & Pone, D. (2011). Permeability

evolution of fluid-infiltrated coal containing discrete fractures. International Journal

of Coal Geology, 85(2), 202–211. http://doi.org/10.1016/j.coal.2010.10.006

Javadpour, F. (2009). Nanopores and apparent permeability of gas flow in mudrocks

(shales and siltstone). Journal of Canadian Petroleum Technology, 48, 1–6.

Javadpour, F., Fisher, D., & Unsworth, M. (2007). Nanoscale gas flow in shale gas

sediments. Journal of Canadian Petroleum Technology, 46(10).

http://doi.org/10.2118/07-10-06

Jones, S. (1997). A technique for faster pulse-decay permeability measurements in tight

rocks. SPE Formation Evaluation, (November 1994), 25–28.

Kamath, J., Boyer, R., & Nakagawa, F. (1992). Characterization of core-scale

heterogeneities using laboratory pressure transients. SPE Formation Evaluation,

(September), 219–227.

Karniadakis, G., Beskok, A., & Narayan, A. (2005). Microflows and Nanoflows.

http://doi.org/10.1007/0-387-28676-4

Klinkenberg, L. J. (1941). The permeability of porous media to liquids and gases.

Drilling and Production Practice, (2), 57–73.

Page 49: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

34

http://doi.org/10.5510/OGP20120200114

Kuila, U., Prasad, M., & Kazemi, H. (2013). Assessing Knudsen flow in gas-flow models

of shale reservoirs (pp. 23–28).

Kumar, A. (2007). Methane Diffusion Characteristics of Illinois Coals.

Kumar, H., Elsworth, D., Liu, J., Pone, D., & Mathews, J. P. (2012). Optimizing

enhanced coalbed methane recovery for unhindered production and CO2 injectivity.

International Journal of Greenhouse Gas Control, 11, 86–97.

http://doi.org/10.1016/j.ijggc.2012.07.028

Kumar, H., Elsworth, D., Mathews, J. P., Liu, J., & Pone, D. (2014). Effect of CO2

injection on heterogeneously permeable coalbed reservoirs. Fuel, 135, 509–521.

http://doi.org/10.1016/j.fuel.2014.07.002

Kumar, H., Elsworth, D., Mathews, J. P., & Marone, C. (2015). Permeability evolution in

sorbing media: analogies between organic-rich shale and coal. Geofluids.

http://doi.org/10.1111/gfl.12135

Kwon, O., Kronenberg, A. K., Gangi, A. F., & Johnson, B. (2001). Permeability of

Wilcox shale and its effective pressure law. Journal of Geophysical Research, 106,

19339. http://doi.org/10.1029/2001JB000273

Langmuir, I. (1918). The adsorption of gases on plane surfaces of glass, mica and

platinum. Journal of the American Chemical Society, 345(1914).

Levine, J. (1996). Model study of the influence of matrix shrinkage on absolute

permeability of coal bed reservoirs. Geological Society, London, Special

Publications.

Li, J., Liu, D., Lu, S., Yao, Y., & Xue, H. (2015). Evaluation and modeling of the CO2

permeability variation by coupling effective pore size evolution in anthracite coal.

Energy & Fuels. http://doi.org/10.1021/ef502674y

Li, J., Liu, D., Yao, Y., Cai, Y., & Chen, Y. (2013). Evaluation and modeling of gas

permeability changes in anthracite coals. Fuel, 111, 606–612.

http://doi.org/10.1016/j.fuel.2013.03.063

Li, Y., Li, X., Shi, J., Wang, H., Wu, L., Teng, S., & Corp, S. (2014). A Nano-Pore Scale

Gas Flow Model for Shale Gas Reservoir Flow characteristic of shale gas reservoir

in various pole scales and molecule scales, (June), 9–11.

Liu, J., Chen, Z., Elsworth, D., Miao, X., & Mao, X. (2010). Linking gas-sorption

induced changes in coal permeability to directional strains through a modulus

reduction ratio. International Journal of Coal Geology, 83(1), 21–30.

http://doi.org/10.1016/j.coal.2010.04.006

Liu, S. (2012). Estimation of different coal compressibilities of coalbed methane

reservoirs under replicated in situ condition.

Liu, S., & Harpalani, S. (2013a). A new theoretical approach to model sorption-induced

coal shrinkage or swelling. AAPG Bulletin, 97(7), 1033–1049.

http://doi.org/10.1306/12181212061

Page 50: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

35

Liu, S., & Harpalani, S. (2013b). Permeability prediction of coalbed methane reservoirs

during primary depletion. International Journal of Coal Geology, 113, 1–10.

http://doi.org/10.1016/j.coal.2013.03.010

Liu, S., & Harpalani, S. (2014a). Compressibility of sorptive porous media: Part 1.

Background and theory. AAPG Bulletin, c, 1–25.

Liu, S., & Harpalani, S. (2014b). Compressibility of Sorptive Porous Media: Part I –

Background and Theory. AAPG Bulletin, 98(9), 1761–1772.

Liu, S., & Harpalani, S. (2014c). Compressibility of Sorptive Porous Media: Part II –

Experimental Study on Coal. AAPG Bulletin, 98(9), 1773–1788.

Liu, S., & Harpalani, S. (2014d). Determination of the effective stress law for

deformation in coalbed methane reservoirs. Rock Mechanics and Rock Engineering,

47(5), 1809–1820. http://doi.org/10.1007/s00603-013-0492-6

Liu, S., & Harpalani, S. (2014e). Evaluation of in situ stress changes with gas depletion

of coalbed methane reservoirs. Journal of Geophysical Research: Solid Earth,

119(8), 6263–6276. http://doi.org/10.1002/2014JB011228.

Liu, S., Harpalani, S., & Pillalamarry, M. (2012). Laboratory measurement and modeling

of coal permeability with continued methane production: Part 2 – Modeling results.

Fuel, 94, 117–124. http://doi.org/10.1016/j.fuel.2011.10.053

Lorenz, J. C., Teufel, L. W., & Warpinski, N. R. (1991). Regional fractures; I, A

mechanism for the formation of regional fractures at depth in flat-lying reservoirs.

AAPG Bulletin , 75(11), 1714–1737.

Luffel, D., Hopkins, C., & Jr, P. S. (1993). Matrix permeability measurement of gas

productive shales. In SPE Annual Technical Conference and Exhibition, Houston,

Texas, 2-6 Oct, 1993.

Ma, Q., Harpalani, S., & Liu, S. (2011). A simplified permeability model for coalbed

methane reservoirs based on matchstick strain and constant volume theory.

International Journal of Coal Geology, 85(1), 43–48.

http://doi.org/10.1016/j.coal.2010.09.007

Malkovsky, V. I., Zharikov, a. V., & Shmonov, V. M. (2009). New methods for

measuring the permeability of rock samples for a single-phase fluid. Izvestiya,

Physics of the Solid Earth, 45(2), 89–100.

http://doi.org/10.1134/S1069351309020013

Markowski, A. K. (2014). Pennsylvania Coalbed Methane Update, 80353.

Markowski, A., Resources, N., & Virginia, W. (2014). Coalbed methane in

Pennsylvania.

Mavor, M. J., Owen, L. B., & Pratt, T. J. (1990). Measurement and Evaluation of Coal

Sorption Isotherm Data. Proceedings of SPE Annual Technical Conference and

Exhibition, (1). http://doi.org/10.2523/20728-MS

Mazumder, S., & Wolf, K. (2008). Differential swelling and permeability change of coal

in response to CO 2 injection for ECBM. International Journal of Coal Geology.

Page 51: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

36

Mckernan, R. E., Rutter, E. H., Mecklenburgh, J., & Taylor, K. G. (2014). Influence of

Effective Pressure on Mudstone Matrix Permeability: Implications for Shale Gas

Production. In SPE/EAGE European Unconventional Conference and Exhibition,

Vienna, Austria, 25-27 Feb, 2014 (p. 13). Society of Petroleum Engineers.

http://doi.org/10.2118/167762-MS

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1–Laboratory results. Fuel,

94, 110–116.

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1 – Laboratory results. Fuel,

94, 110–116. http://doi.org/10.1016/j.fuel.2011.10.052

Okotie, V. U., & Moore, R. L. (2011). Well-production challenges and solutions in a

mature, very-low-pressure coalbed-methane reservoir. SPE Production and

Operations, 26(2), 149–161. http://doi.org/10.2118/107705-MS.

Palmer, I., & Mansoori, J. (1998). How permeability depends on stress and pore pressure

in coalbeds: a new model. SPE Reservoir Evaluation and Engineering, SPE

52607(December), 539–544.

Palmer, S. I., & Mansoori, J. (1998). How permeability depends on stress and pore

pressure in coalbeds: a new model. Soc Petrol Eng Res Eng, (December), 539–544.

Pan, Z., & Connell, L. D. (2007). A theoretical model for gas adsorption-induced coal

swelling. International Journal of Coal Geology, 69(4), 243–252.

http://doi.org/10.1016/j.coal.2006.04.006

Pillalamarry, M., Harpalani, S., & Liu, S. (2011). Gas diffusion behavior of coal and its

impact on production from coalbed methane reservoirs. International Journal of

Coal Geology, 86(4), 342–348. http://doi.org/10.1016/j.coal.2011.03.007

Randolph, P. L., Soeder.D.J., & Chowdiah, P. (1984). Porosity and Permeability of Tight

Sands. In SPE/DOE/GRI Unconventional Gas Recovery Symposium held in

Pittsburgh, PA, May 13-15. Dallas.

Reif.F. (1965). Fundamentals of statistical and thermal physics.

Rexer, T. F. T., Benham, M. J., Aplin, A. C., & Thomas, K. M. (2013). Methane

Adsorption on Shale under Simulated Geological Temperature and Pressure

Conditions. Energy & Fuels, 27(6), 3099–3109. http://doi.org/10.1021/ef400381v

Rodrigues, C. F., & Lemos De Sousa, M. J. (2002). The measurement of coal porosity

with different gases. International Journal of Coal Geology, 48(3–4), 245–251.

http://doi.org/10.1016/S0166-5162(01)00061-1

Roy, S., Raju, R., Chuang, H. F., Cruden, B. a., & Meyyappan, M. (2003). Modeling gas

flow through microchannels and nanopores. Journal of Applied Physics, 93(8),

4870–4879. http://doi.org/10.1063/1.1559936

Ruckenstein, E. (1971). Sorption by solids with bidisperse pore structures. Chemical

Engineering …, 26, 1305–1318.

Page 52: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

37

Rydzy, M. B., Patino, J., Elmetni, N., & Appel, M. (2016). Stressed Permeability in

Shales : Effects of Matrix Compressibility and Fractures - A Step towards

Measuring Matrix Permeability in Fractured Shale Samples, 1–12.

Sakhaee-Pour, A., & Bryant, S. (2012). Gas Permeability of Shale. SPE Reservoir

Evaluation & Engineering, 15, 401–409.

Seidle, J. P., Jeansonne, M. W., & Erickson, D. J. (1992). Application of matchstick

geometry to stress dependent permeability in coals. SPE Rocky Mountain Regional

Meeting, 433–444. http://doi.org/10.2523/24361-MS

Shabro, V. (2012). Forecasting Gas Production in Organic Shale with the Combined

Numerical Simulation of Gas Diffusion in Kerogen, Langmuir Desorption from

Kerogen Surfaces, and Advection in Nanopores. SPE Annual Technical ….

http://doi.org/10.2118/159250-MS

Shi, J.-Q., Pan, Z., & Durucan, S. (2014). Analytical models for coal permeability

changes during coalbed methane recovery: Model comparison and performance

evaluation. International Journal of Coal Geology, 136, 17–24.

http://doi.org/10.1016/j.coal.2014.10.004

Shi, J., & Durucan, S. (2005). A model for changes in coalbed permeability during

primary and enhanced methane recovery. SPE Reservoir Evaluation & Engineering,

8(4), 291–299.

Shi, J. Q., & Durucan, S. (2003). A bidisperse pore diffusion model for methane

displacement desorption in coal by CO 2 injection☆. Fuel, 82(10), 1219–1229.

http://doi.org/10.1016/S0016-2361(03)00010-3

Shi, J. Q., & Durucan, S. (2004). Drawdown Induced Changes in Permeability of

Coalbeds: A New Interpretation of the Reservoir Response to Primary Recovery.

Transport in Porous Media, 56(1), 1–16.

http://doi.org/10.1023/B:TIPM.0000018398.19928.5a

Shi, J. Q., & Durucan, S. (2014). Modelling laboratory horizontal stress and coal

permeability data using S&D permeability model. International Journal of Coal

Geology, 131, 172–176. http://doi.org/10.1016/j.coal.2014.06.014

Singh, H., & Javadpour, F. (2013). Nonempirical Apparent Permeability of Shale. In

Unconventional Resources Technology Conference (Vol. 2, p. 16). Denvor,

Colorado.

Singh, H., & Javadpour, F. (2015). Langmuir slip-Langmuir sorption permeability model

of shale. FUEL, (October). http://doi.org/10.1016/j.fuel.2015.09.073

Smith, D., & Williams, F. (1984). Diffusional effects in the recovery of methane from

coalbeds. SPEJ, Soc. Pet. Eng. J.;(United States), (October), 529–535.

Soeder, D. (1988). Porosity and permeability of eastern Devonian gas shale. SPE

Formation Evaluation, 3, 116–124.

Sudibandriyo, M., Pan, Z., Fitzgerald, J. E., Jr, R. L. R., & Gasem, K. A. M. (2003).

Adsorption of Methane, Nitrogen, Carbon Dioxide, and Their Binary Mixtures on

Page 53: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

38

Dry Activated Carbon at 318.2 K and Pressures up to 13.6 MPa. Lan, 19(8), 5323–

5331.

Suggested Methods for Uniaxial‐Strain Compressibility Testing for Reservoir

Geomechanics. (2015), c(April), 1–29.

Swami, V. (2012). Shale Gas Reservoir Modeling: From Nanopores to Laboratory. SPE

Annual Technical Conference and Exhibition, (October), 8–10.

Swami, V., Clarkson, C., & Settari, A. (2012). Non-Darcy Flow in Shale Nanopores: Do

We Have a Final Answer? SPE Canadian Unconventional …, (November), 1–17.

Swami, V., & Settari, A. (2012). A pore scale gas flow model for shale gas reservoir. In

SPE Americas Unconventional Resources Conference held in Pittsburgh (pp. 5–7).

Terzaghi, K. (1943). Theoretical Soil Mechanics. John Wiley & Sons, Inc.

http://doi.org/10.1680/geot.1964.14.1.1

Teufel, L. W., Rhett, D. W., & Farrell, H. E. (1991). Effect of Reservoir Depletion And

Pore Pressure Drawdown On In Situ Stress And Deformation In the Ekofisk Field,

North Sea. The 32nd U.S. Symposium on Rock Mechanics (USRMS), 63–72.

U S Energy Information Administration. (2011). Emissions of Greenhouse Gases in the

United States 2009, (March).

Walls, J., Diaz, E., & Cavanaugh, T. (2012). Shale Reservoir Properties from Digital

Rock Physics. Proceedings of SPE/EAGE European Unconventional Resources

Conference and Exhibition, 1–9. http://doi.org/10.2118/152752-MS

Wang, G., Ren, T., Wang, K., & Zhou, A. (2014). Improved apparent permeability

models of gas flow in coal with Klinkenberg effect. Fuel, 128, 53–61.

http://doi.org/10.1016/j.fuel.2014.02.066

Wang, S., Elsworth, D., & Liu, J. (2011). Permeability evolution in fractured coal: the

roles of fracture geometry and water-content. International Journal of Coal

Geology, 87(1), 13–25. http://doi.org/10.1016/j.coal.2011.04.009

Wang, S., Elsworth, D., & Liu, J. (2012). A mechanistic model for permeability evolution

in fractured sorbing media. Journal of Geophysical Research, 117(B6).

Wang, S., Elsworth, D., & Liu, J. (2012a). A mechanistic model for permeability

evolution in fractured sorbing media. Journal of Geophysical Research: Solid Earth

( …, 117(B6), B06205. http://doi.org/10.1029/2011JB008855

Wang, S., Elsworth, D., & Liu, J. (2012b). Permeability evolution during progressive

deformation of intact coal and implications for instability in underground coal

seams. International Journal of Rock Mechanics and Mining Sciences, 1–12.

Wang, Y., & Liu, S. (2016). Estimation of Pressure-Dependent Diffusive Permeability of

Coal Using Methane Diffusion Coefficient: Laboratory Measurements and

Modeling. Energy and Fuels, 30(11), 8968–8976.

http://doi.org/10.1021/acs.energyfuels.6b01480

Wang, Y., Liu, S., & Elsworth, D. (2015). Laboratory investigations of gas flow

Page 54: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

39

behaviors in tight anthracite and evaluation of different pulse-decay methods on

permeability estimation. International Journal of Coal Geology, 149, 118–128.

http://doi.org/10.1016/j.coal.2015.07.009

Wu, K., Li, X., Wang, C., Chen, Z., & Yu, W. (2015). A Model for Gas Transport in

Microfractures of Shale and Tight Gas Reservoirs. AIChE Journal, 61(6).

http://doi.org/10.1002/aic

Wu, K., Li, X., Wang, C., Yu, W., & Chen, Z. (2014). Apparent Permeability for Gas

Flow in Shale Reservoirs Coupling Effects of Gas Diffusion and Desorption.

Proceedings of the 2nd Unconventional Resources Technology Conference, 1–18.

http://doi.org/10.15530/urtec-2014-1921039

Xu, H., Tang, D., Zhao, J., Li, S., & Tao, S. (2015). A new laboratory method for

accurate measurement of the methane diffusion coefficient and its influencing

factors in the coal matrix. Fuel, 158, 239–247.

http://doi.org/10.1016/j.fuel.2015.05.046

Yi, J., Akkutlu, I., & Deutsch, C. (2008). Gas Transport in Bidisperse Coal Particles:

Investigation for an Effective Diffusion Coefficient in Coalbeds. Journal of

Canadian Petroleum Technology, 47(10). http://doi.org/10.2118/08-10-20

Yin, G., Jiang, C., Wang, J. G., & Xu, J. (2013). Combined Effect of Stress, Pore

Pressure and Temperature on Methane Permeability in Anthracite Coal: An

Experimental Study. Transport in Porous Media, 100(1), 1–16.

http://doi.org/10.1007/s11242-013-0202-6

Yuan, W., Pan, Z., Li, X., Yang, Y., Zhao, C., Connell, L. D., … He, J. (2014).

Experimental study and modelling of methane adsorption and diffusion in shale.

Fuel, 117, 509–519. http://doi.org/10.1016/j.fuel.2013.09.046

Zhang, T., Ellis, G. S., Ruppel, S. C., Milliken, K., & Yang, R. (2012). Effect of organic-

matter type and thermal maturity on methane adsorption in shale-gas systems.

Organic Geochemistry, 47, 120–131.

http://doi.org/10.1016/j.orggeochem.2012.03.012

Zoback, M. D., & Byerlee, J. D. (1975). Permeability and Effective Stress. Aapg Bulletin-

American Association of Petroleum Geologists, 59(1), 154–158.

http://doi.org/10.1306/83D91C40-16C7-11D7-8645000102C1865D

Page 55: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

40

Appendix

Figure 1 Annual CBM production in USA from1989 to 2013 (EIA, 2014)

Figure 2 Photograph of cylindrical Hazelton anthracite core samples.

Page 56: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

41

Figure 3 Picture of the pulse-decay experimental setup for permeability evolution test. V1 is the

valve controlling inlet gas flow from upstream, V2 is the valve controlling outlet gas flow to

downstream and V3/V4 is the valve controlling the confining/axial stress applied on the sample.

Figure 4 Rubber jacket deformation evolution with confining stress (Linear Relationship).

0

0.001

0.002

0.003

0.004

0 500 1000 1500 2000 2500 3000

Th

ick

nes

s D

eform

ati

on

(m

m)

Confining Stress (psi)

Page 57: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

42

Figure 5 Comparison of pulse-decay pressure responses for helium and CO2 injections. The

pressure equilibrium time for helium is at least 30 hours less than that of CO2 at each pressure

step. At high pressure helium pressure can get equilibrium in a relatively short time while the

CO2 pressure yet have a longer.

Page 58: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

43

Figure 6 Carbon dioxide pulse-decay pressure curves at three different pore pressure under

constant stress condition in gas depletion process.

Figure 7 Helium permeability profiles under constant stress condition, obtained by Brace’s and

Dicker & Smits’s methods respectively.

0

0.002

0.004

0.006

0.008

0.01

0.012

0 200 400 600 800 1000

Per

mea

bil

ity (

mD

)

Pore Pressure (psi)

Brace et al., 1968

Dicker & Smits, 1988

Page 59: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

44

Figure 8 Helium permeability profiles under uniaxial strain condition, obtained by Brace’s and

Dicker & Smits’s methods respectively.

0

0.002

0.004

0.006

0.008

0.01

0.012

0 200 400 600 800 1000

Per

mea

bil

ity (

mD

)

Pore Pressure (psi)

Brace et al., 1968

Dicker & Smits, 1988

Page 60: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

45

Figure 9 Permeability increments of Dicker & Smits's method comparing to Brace's method at

constant pressure at 350 psi.

Figure 10 Horizontal stress variation for helium depletion under uniaxial strain condition.

0

5

10

15

0 5 10 15 20 25 30 35

Per

mea

bil

ity I

ncr

emen

t (%

)

Porosity (%)

0

300

600

900

1200

1500

1800

0 200 400 600 800 1000

Con

fin

ing S

tres

s (p

si)

Pore Pressure (psi)

Page 61: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

46

Figure 11 Change in effective horizontal stress with helium depletion under two different stress-

strain conditions: uniaxial strain and constant stress.

Figure 12 Permeability evolution with helium depletion under both constant stress and uniaxial

strain condition using Dicker & Smits’s method.

0

200

400

600

800

1000

1200

0 200 400 600 800 1000

Eff

ecti

ve

Hori

zon

tal

Str

ess

(psi

)

Pore Pressure (psi)

Stress-controlled Condition

Uniaxial-strain Condition

0

0.002

0.004

0.006

0.008

0.01

0.012

0 200 400 600 800 1000

Per

mea

bil

ity (

mD

)

Pore Pressure (psi)

Stress-controlled

Condition

Uniaxial-strain

Condition

Page 62: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

47

Figure 13 Comparison between three different pulse-decay methods with effective stress.

Figure 14 Experimental results comparison with anthracite CO2 permeability data by Wang et al.

2011.

0

200

400

600

800

1000

1200

0

0.0001

0.0002

0.0003

0.0004

0 200 400 600 800 1000

Effectiv

e Str

ess (psi)P

erm

eab

ilit

y (

mD

)

Pore Pressure (psi)

Brace et al., 1968

Dicker & Smits, 1988

Cui et al., 2009

Effective Stress

0

0.0001

0.0002

0.0003

0.0004

0 200 400 600 800 1000

Per

mea

bil

ity (

mD

)

Pore Pressure (psi)

Experimental Data

Wang et al. 2011

Page 63: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

48

Figure 15 Comparison between intrinsic permeability and apparent permeability obtained by Cui

et al.’s methods.

Table 1 Comparison and evaluation of different pulse-decay methods Method Description Comments

Brace et al. (1968) Original pulse-decay method Assumed no compressive gas

storage and pure Darcy flow

Hsieh et al. (1981) Presented a general analytical solution

for compressive storage of sample

Dicker & Smits (1988) Applied compressive storage effect into Without considering the effect

pulse-decay measurement of gas adsorption/desorption

Kamath et al. (1992) Pulse-decay to interpret the core’s heterogeneity

Jones (1997) Simplified the compressive storage factor

Cui et al. (2009) Added adsorption component in the Considering both compressive

compressive factor storage and adsorption effect

Wang et al. (2011) Comprehensive pulse-decay test on Simply used the original

Anthracite coals pulse-decay method

Mckernan et al. (2014) Used oscillating pore pressure method to

get pressure decay

Table 2 Proximate analysis of Pennsylvania anthracite coal sample

Proximate Analysis

Fixed carbon Moisture Ash content Volatile matter

0

0.0001

0.0002

0.0003

0.0004

0 200 400 600 800 1000

Per

mea

bil

ity (

mD

)

Pore Pressure (psi)

Cui et al., 2009_intrinsic permeability

Cui et al., 2009_apparent permeability

Page 64: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

49

78.35% 1.59% 10.79% 9.27%

Table 3 Experimental data of helium permeability (mD) under stress-controlled and uniaxial-strain

conditions

Pore Pressure (psi)

Boundary Condition 150 230 350 440 580 700

Stress-controlled

Uniaxial-strain

0.000449 0.000442 0.000486 0.000614 0.00204 0.00320

0.000469 0.000461 0.000507 0.000640 0.00213 0.00334

Table 4 Experimental data of CO2 permeability (mD) by using different pulse-decay approaches

Pore Pressure (psi)

Pulse-decay Method 30 140 250 350 550 875

Brace et al., 1968

Dicker & Smits, 1988

Cui et al., 2009

8.36E-05 6.28E-05 3.75E-05 1.91E-05 2.18E-05 2.4E-04

8.40E-05 6.31E-05 3.76E-05 1.92E-05 2.19E-05 2.4E-04

1.05E-04 7.26E-05 4.31E-05 2.12E-05 2.33E-05 2.5E-04

Page 65: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

50

Chapter 3 Estimation of Pressure-dependent Diffusive Permeability of

Coal Using Methane Diffusion Coefficient: Laboratory Measurements

and Modeling

Abstract

Gas diffusion process in coal is critical for the prediction of coalbed methane

production, especially for the late-time CBM reservoir when both gas pressure and

permeability is relative low. Using only Darcy permeability to evaluate the quality of gas

transport may not be effective. Diffusive flow can be dominant flow at low reservoir

pressure. In this work, methane diffusion coefficient was measured on pulverized San Juan

sub-bituminous and Pittsburgh bituminous coal samples using classic unipore model and

particle method. The diffusion coefficient results showed a negative correlation with

pressure which has been reported before. And the significance of diffusion flow is strongly

related to the rate of methane ad-/desorption process, severer at low pressure range (< 2

MPa/280 psi). The measured diffusion coefficient can be converted to the equivalent

permeability. This equivalent permeability can be considered as the contribution of

diffusion flow, in terms of Darcy permeability, used to evaluate total gas flow at late

production decline stage when matrix flow dominates. As expected, this diffusive

permeability was found to be much lower than the fracture/cleat permeability. An

increasing trend at low pressure due to pressure drop was obtained and coincides with

sorbing gas permeability behavior when the pressure is extremely low.

Page 66: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

51

1. Introduction

Coalbed methane (CBM) is a clean-burning energy source well suited as a fuel for

production of electricity, residential and commercial heating, and as a vehicle fuel. CBM

currently supplies approximately eight percent of the nation’s natural gas production, and

is an important portion of the nation’s energy mix. From an environmental standpoint,

methane emission from coal mines poses an environmental risk since methane is a potent

greenhouse gas, second only to CO2(U S Energy Information Administration, 2011).

The San Juan basin is one of the oldest and largest CBM productive areas in North

America. North Appalachian basin has abundant CBM resource.(A. Markowski et al.,

2014) With a long history of production, some gas wells in San Juan and north Appalachian

coal fields are now in their mature stage of reservoir-pressure depletion.2,3 For mature CBM

wells, the diffusion flow dominates the gas transport since the permeability is orders of

magnitude higher than the matrix diffusive gas flow at very low reservoir pressure (< 50

psi for some San Juan old wells). Diffusive flow plays an increasingly important role as

the reservoir pressure depletion. Figure 16 shows a typical CBM well production behavior.

The water production declines rapidly until the gas reaches the peak production. The

dewatering process can take a few months or years and then the gas production reaches its

stable production stage associated with limited amount of water production.(Aminian &

Ameri, 2004) After the stable production stage, gas production starts to decline which is

termed as decline production stage. Then the gas production rate becomes relatively flat

and the flat production tail may last for years or even decades. In the flat tail production

stage, the reservoir permeability is high due to the matrix shrinkage induced cleat/fracture

opening.5,6 In terms of gas production, the coal permeability may not be the bottle-neck for

Page 67: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

52

the mature CBM wells as it did in early stage, because the gas mass influx from matrix to

cleats, controlled by diffusion process, starts to control the overall gas production in those

mature wells. As shown in Figure 16, the orange and purple dash production profiles

represent the high and low diffusion potential coal at the late production stage. A detailed

understanding of the diffusive flow that is controlled by diffusion coefficient will help the

gas operators to decide: what is going to be the abandonment pressure of the well; when

does the well need a cleanup and/or restimulation, and what is going to be the well life?

In this study, the diffusion coefficients were first experimentally measured by the

classic particle method and then the data was modeled by unipore diffusion model for both

San Juan and Pittsburgh coal samples. The discussion towards the behavior of diffusion

coefficient will provide the information about how gas diffusion process is influenced by

pressure and adsorption effect. The estimated diffusion coefficient was transformed to

format of permeability to quantitatively evaluate the gas transport intensity in coal at low

pressure in late production stage.

2. Background and literature review

2.1 Gas diffusion in coal

It is widely known that coal matrix is a dual porosity system, and methane transport

mechanism in CBM can be differentiated as dewatering, gas desorption, diffusion and

Darcy viscous flow.7,8 The tight structure of coal introduces various flow dynamics

mechanism which results in the challenge of CBM reservoir evaluation. The rate of gas

flow between the bulk solid matrix and its surface is very slow(Swami & Settari, 2012)

and was believed to be less important from the point of view of well production. Diffusion,

Page 68: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

53

however, plays an important role and diffusivity is a critical parameter of tight reservoir

rocks, especially at low pressure (< 50-100 psi).10,11 The diffusion process is described by

Fick’s Second Law and is driven by gas concentration gradient. When Knudsen number

(𝐾𝑛), defined as the ratio of the molecular mean free path length to characteristic diameter

of pore, is relatively high ( 𝐾𝑛 > 10), the diffusion can dominate the overall mass

transport.(Karniadakis, Beskok, & Narayan, 2005) Fick’s Law is used to relate the

diffusive mass transfer to concentration gradient by assuming that the mass flowrate across

a surface is proportional to the concentration gradient across the surface, area of the surface,

and diffusion coefficient for the solid/gas medium.(A. Kumar, 2007) Based on Fick’s Law,

the diffusion coefficient is introduced as a factor indicating the significance of the diffusion

process. The diffusion in coal matrix depends upon the matrix structure and pressure.9,14,15

Through studying the diffusion rates of different ranks of coals, Clarkson and Bustin(C.

Clarkson & Bustin, 1999) observed that the difference on adsorption capacity between coal

ranks are mainly due to different proportions of macropores, mesopores and micropores.

Figure 17 shows the gas flow dynamics in coal as a sequential manner. Gas molecules

desorb from the micropore and then diffuse towards the cleat/fracture system. And then

gas flows to the wellbore by pressure driven Dacian flow. Micropores and mesopores in

coal matrix are residential area for gas molecules to store as adsorbed phase and also are

pathways for gas diffusion; macropores and microfractures make the way for gas viscous

flow and laminar water flow.(Cai et al., 2013) Gas molecules desorb from coal matrix

interface, creating a gas concentration gradient between small-scale pores and large-scale

pores.(A. Kumar, 2007)

Page 69: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

54

Potential gas diffusion mechanisms involve molecular diffusion, Knudsen diffusion,

and surface diffusion. When the mean free path of the gas molecules is greater than the

molecular diameter, or when the pressures are very low, Knudsen diffusion takes place,

and gas molecules flow from higher to lower gas concentration.(He, Lv, & Dickerson,

2014; A. Kumar, 2007) Surface diffusion of gas occurs when adsorbed gas molecules move

along the micropore surface like a liquid. At room temperature, surface diffusion is much

less significant than Knudsen diffusion.(Pillalamarry et al., 2011) Typically, this is ignored

for gas deliverability prediction for CBM reservoirs. Molecular/bulk diffusion occurs at

higher pressures and/or when pore diameter is larger than the mean free path of gas

molecules.(Dutta, 2009) Based on recent research findings, Knudsen diffusion can be

reasonably considered as the main diffusion mechanism controlling the gas flow in

unconventional reservoir rocks.14,15,20

2.2 Diffusion coefficient measurement techniques

In order to quantify the diffusive flow, a laboratory approach “Particle Method” was

applied on coal particles.(Pillalamarry et al., 2011) This method is the most commonly

used technique to determine the diffusion coefficient of coal.(A. Kumar, 2007) Assuming

the concentration at the external surface is held constant, the rate of uptake by a spherical

sorbent particle has been given by.(Smith & Williams, 1984) This is commonly referred to

as the unipore model, and was used to calculate the diffusion coefficient in this study. This

type of rate model is appropriate when solid diffusion or diffusion in pores of uniform size

is involved. On the other hand, in some solids with bi-disperse pore size distributions,

diffusion and sorption occur simultaneously in both macropores and

Page 70: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

55

micropores.(Ruckenstein, 1971) Diffusive mass transfer processes such as between the

interior of the grains and the surrounding fluid are of particular importance in many porous

samples with a bi-disperse pore size distribution.(Drazer, Chertcoff, Bruno, & Rosen,

1999) Ruckenstein et al.(Ruckenstein, 1971) developed the bidisperse model to account

for both macropore and micropore diffusion in the matrix using a simplified approach, in

which the adsorbent is spherical and contains microporous spherical particles separated by

inter-particle macropores.

Both unipore and bi-disperse model was reported able to well describe the methane

diffusivity in coal gas production case by case.(Pillalamarry et al., 2011; J.Q. Shi &

Durucan, 2003; Smith & Williams, 1984; Xu, Tang, Zhao, Li, & Tao, 2015; Yuan et al.,

2014) However, the bi-disperse model is considered unclear with the determination of

some arbitrary modeling parameters of coal.(C. R. Clarkson & Bustin, 1999) For

simplicity, the unipore model is adequate for CBM diffusion modeling as a classic

analytical approach.

2.3 Gas diffusive permeability modeling

For coals, diffusion can be the dominant transport mechanism rather than Darcy flow

at the late-time production stage. In CBM production, the permeability of viscous flow

path, including macropores and microfracture system, can potentially be enhanced quite

significantly due to coal matrix shrinkage caused by sorption-induced deformation.(S. Liu

& Harpalani, 2014d) With continuous depletion, the dominant factor can gradually change

from Darcy flow to diffusion flow because the permeability is much larger compared to

the diffusive mass influx as shown in Figure 17. When overall gas transport in coal is

Page 71: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

56

primarily controlled by diffusion flow, the diffusion coefficient should be the used to

predict the gas production performance rather than Darcy permeability. For current

reservoir simulators including ARI COMET3 reservoir simulator, CMG-GEM simulator,

IHS-CBM simulator, however, permeability is still the key input parameter serving as the

most important variable for the prediction of gas production. Technically, the dynamic

diffusion coefficient cannot be directly input to the simulators for the late time production

forecasting. Thus, it is critical to either program the CBM reservoir simulator in order to

properly weight the contribution of diffusion flow by simply treating diffusion coefficient

as a key parameter, or convert diffusion coefficient into the form of Darcy permeability

that can be easily adopted in the current simulators. Obviously, the diffusion coefficient to

permeability conversion is a painless pathway to properly adopt the late time diffusion

controlled flow since no major modifications in simulators are needed. Gas diffusion

coefficient can be transferred to the form of permeability based on the mass conservation

law and the relationship has been theoretically proposed in Cui et al..(X. Cui et al., 2009)

Assuming that only Darcy’s flow regime is dominating, gas transport in porous rocks with

slab, cylindrical or spherical shape can be described by a one-dimension mass balance

equation shown as:

𝜑𝜕𝜌

𝜕𝑡+ (1 − 𝜑)

𝜕𝑞

𝜕𝑡=

1

𝑟𝑚

𝜕

𝜕𝑟(𝑟𝑚 𝜌𝑘

𝜇

𝜕𝑝

𝜕𝑟) (14)

where, t is time, 𝜌 is gas density, q is adsorbate density per unit sample volume, 𝜑 is

porosity of rock sample, p is pressure, k is Darcy permeability, 𝜇 is gas viscosity, r is pore

radius, and m= 0, 1 and 2 represents slab, cylindrical and spherical shapes, respectively. If

Page 72: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

57

we assume the gas transport is dominated by gas diffusion through micropores, governed

by Fick’s law, the above equation can become:

𝜑𝜕𝜌

𝜕𝑡+ (1 − 𝜑)

𝜕𝑞

𝜕𝑡=

1

𝑟𝑚

𝜕

𝜕𝑟(𝑟𝑚𝜑𝐷

𝜕𝜌

𝜕𝑟) 𝑏 =

𝑉𝑝(1+𝜙𝑎𝜙

)

𝑉𝑑 (15)

where, D is diffusion coefficient.

By comparing equation (1) and (2) with applying chain rule, the permeability can be

calculated by:

𝑘 = 𝐷𝜙𝜇𝐶𝑔𝑏 =𝑉𝑝(1+

𝜙𝑎𝜙

)

𝑉𝑑 (16)

where, 𝑐𝑔 is gas isothermal compressibility. An equivalent gas permeability can thus be

estimated if the gas diffusion coefficient can be determined, and vice versa. This equation

indicates the diffusion-based permeability is a function of diffusion coefficient, gas density

and viscosity, rather than physical properties of rock.(X. Cui et al., 2009) Therefore, the

estimated permeability in equation (3) is an apparent permeability which is not purely

determined by the rock structure as the intrinsic permeability does. We term this estimated

permeability as diffusion apparent permeability in the study. The diffusion apparent

permeability can be considered a representation of diffusion flow, in terms of Darcy

permeability, contributes to matrix flow at late time production stage. This helps us to get

a better and accurate late time production forecasting for mature CMB wells.

3. Experimental work

3.1 Sample procurement and preparation

The coal blocks were sub-bituminous coal from San Juan Basin and bituminous coal

from North Appalachian Basin. The coal samples were pulverized into particles at an

Page 73: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

58

average size of 0.5 mm (35 mesh) and dried in the oven at 375K for 24 hours. The coal

proximate analysis was conducted by the Standard Laboratories, Inc. and the results are

listed in Table 5. These pulverized samples were used to conduct adsorption and diffusion

coefficient measurements. By crushing coal into small particles, we eliminated the micro-

fractures in the tested coals to ensure micro-scale flow (diffusion) dominates instead of

Darcy flow. This is the key of “particle method” and can guarantee a faster and more

effective adsorption measurement at the same time.

3.2 Experimental setup and procedure

The sorption capacity and diffusion coefficient are simultaneously estimated using

volumetric gas adsorption apparatus shown in Figure 18 and a schematic view in Figure

19. The experimental system was built with high strength stainless steel with Swagelok

tubing and fittings. Both sample and reference volumes were measured by helium

expansion. The volumes of reference and sample cells were 158.95 mL and 136.32 mL.

The sample cell was connected to the reference cell through a two-way valve. A micro-

filter was install between two cells to prevent movement of coal particles from sample cell

to reference cell due to the sudden pressure shocks. The apparatus was placed in a

programmable temperature-controlled water bath to ensure the constant temperature for

the sorption and diffusion measurements. Both the sample and reference cell pressures

were continuously monitored at each pressure step. Samples were flooded with methane in

a stepwise manner up to ~1300 psi (9 MPa). Equilibration relaxation kinetics were

monitored during the experiment and the data recorded during the experiment will be

pressures with time forming the basis for all the calculations.

Page 74: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

59

3.3 Ad-desorption isotherm measurement and estimation

The real gas equation of state (EOS) was used for the Gibbs adsorption amount

estimation. Through each gas injection step, we were able to obtain a mass balance equation,

and the amount of Gibbs adsorption (excess) can be directly calculated. The moles of

adsorbed gas can be calculated by the following equation:

𝑛𝑚 = 𝑃𝑖𝑛𝑖𝑉𝑅

𝑍1𝑅𝑇−

𝑃𝑝𝑜𝑟𝑒(𝑉𝑅+𝑉𝑣𝑜𝑖𝑑)

𝑍2𝑅𝑇 (17)

where, 𝑛𝑚 is the Gibbs adsorption amount in mole, 𝑃𝑖𝑛𝑖 is the initial injection pressure

before sorption, 𝑃𝑝𝑜𝑟𝑒 is the final equilibrium pressure, 𝑉𝑅 is the volume of reference cell,

𝑉𝑣𝑜𝑖𝑑 is the void volume in the sample cell that is not occupied by solid particles, 𝑍1, 𝑍2

are gas compressibility factor of pressure 𝑃𝑖𝑛𝑖 and 𝑃𝑝𝑜𝑟𝑒, R is the universal gas constant,

and T is temperature.

Gibbs adsorption is good for the convenience of laboratory measuring adsorption

amount through gas injections, but it neglects the volume occupied by the adsorbed phase

in calculating the amount of unadsorbed gas.(Pillalamarry et al., 2011) Based on the Gibbs

adsorption amount (𝑛𝑚), the absolute adsorption at standard condition was estimated by

the following equation(Pillalamarry et al., 2011; Rexer, Benham, Aplin, & Thomas, 2013;

Sudibandriyo, Pan, Fitzgerald, Jr, & Gasem, 2003):

𝑛𝑎𝑏𝑠 = 𝑛𝑚

1−𝜌

𝜌𝑎𝑑𝑠

(18)

where, 𝑛𝑎𝑏𝑠 is the absolute adsorption amount, and 𝜌𝑎𝑑𝑠 is the adsorbed phase density.

Eventually, the absolute adsorption amount will be used to establish the Langmuir

Page 75: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

60

isotherms for adsorption results evaluation. The Langmuir model is mathematically

described as:

𝑉𝑎𝑏𝑠 = 𝑃𝑉𝐿

𝑃+𝑃𝐿 (19)

where, 𝑉𝑎𝑏𝑠 is the absolute adsorption volume, 𝑉𝐿 is the Langmuir Volume representing

the maximum volume that can be adsorbed at infinite pressure, and 𝑃𝐿 is the Langmuir

Pressure at which the absolute adsorption volume is half the Langmuir Volume.

3.4 Diffusion coefficient estimation

The particle method was used to estimate the diffusion coefficient for coal powder

samples. Same as adsorption experiments, the reference cell was subjected to methane

pressure higher than in a sample container. For every step increase in pressure, there was

adsorption of methane happening in the matrix resulting in a gradual decrease of pressure

in the sample container. This pressure decrease due to diffusion was observed in a short

time period and measured very precisely and the pressure decrease profile and time was

used to estimate the coal diffusion coefficient. The pressure versus time data will be used

to calculate diffusion coefficient at the initial stage of each injection step. The diffusion

coefficient measurement was simultaneously conducted with adsorption experiments and

can be used to determine the flow rate of diffusive flow.

Unipore diffusion model, assuming all the pores in the coal matrix are of the same

radius, was used to derive the particle method. The model also assumes a constant gas

concentration at the surface of the spheres throughout the sorption process (Pillalamarry et

al., 2011). And the basis of this method is Fick’s Second Law for spherically symmetric

Page 76: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

61

flow, given as:

𝐷

𝑟2

𝜕

𝜕𝑟(𝑟2 𝜕𝐶

𝜕𝑟) =

𝜕𝐶

𝜕𝑡 (20)

where, C is the adsorbate concentration, and D is the diffusion coefficient

The solution to Equation (7) for a constant surface concentration of the diffusing can

be expressed as follows(C. R. Clarkson & Bustin, 1999; Crank, 1957):

𝑀𝑡

𝑀∞= 1 −

6

𝜋2∑

1

𝑛2 exp (−𝐷𝑛2𝜋2𝑡

𝑟2 ) (21) ∞𝑛=1

where, 𝑀𝑡 is the total mass of the diffusing gas that has adsorbed in time t, and 𝑀∞ is the

total ad-/desorbed mass in infinite time. This solution is described by the total amount of

diffusing substance entering a sphere media, adsorption in coal matrix in this case. This

relationship can also be expressed in terms of adsorption volume (C. R. Clarkson & Bustin,

1999):

𝑉𝑡

𝑉∞= 1 −

6

𝜋2∑

1

𝑛2 exp (−𝐷𝑛2𝜋2𝑡

𝑟2 ) (22) ∞𝑛=1

where, 𝑉𝑡 is the total volume of the diffusing gas that has ad-/desorbed in time t, and 𝑉∞ is

the total ad-/desorbed volume in infinite time. These two equations are commonly referred

to the unipore model equation, and Equation (8) was used to calculate the diffusion

coefficient in this study.

4. Results and Discussion

4.1 Adsorption isotherm results

Experimental data for methane absolute adsorption isotherms for crushed San Juan

and Pittsburgh #8 coals are collated according to the volumetric method by Mavor et

Page 77: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

62

al.( 1990) The temperature was kept constant at 35 C o during the entire experiment. The

Gibbs adsorption and absolute adsorption mass was calculated using Equations (4) and (5),

with adsorbed phase density assumed to be 0.421 g/mL. The results are shown in Figure

20. As expected, the adsorption isotherms for both San Juan and Pittsburgh #8 coal samples

are following the trend of typical Langmuir type adsorption isotherm. The Langmuir

parameters were estimated based on the adsorption isotherm. The Langmuir volume and

pressure are 1.12 mmol/g and 3.14 MPa respectively for Suan Juan coal sample; 0.78

mmol/g and 4.04 MPa for Pittburgh #8 coal sample. Obviously, San Juan coal sample has

a higher adsorption capacity and lower Langmuir pressure than Pittsburgh #8 coal sample,

which indicates the former has better overall adsorption potential than the latter.

4.2 Methane diffusion coefficient results

Methane diffusion coefficients of two tested pulverized coals were estimated by using

Equation (8) and the results are shown in Figure 21 and 22. Both diffusion curves showed

a negative correlation when gas pressure became larger than 2 MPa. There is also a slight

increasing trend of CH4 diffusion plot when pressure is below 2 MPa. The assumption to

explain this phenomenon is that a large quantity of unopened micropores become open

after pressure was introduced for couple of hours. More gas pathways were created for gas

molecules to diffuse in, and so as to the more significant diffusion flow. In the

measurements, the adsorption amount (𝑀𝑡) at time t was used in the calculation. Three time

durations, at 10, 20, and 30 minutes, were collected respectively. The reason why we chose

three different time periods is mainly because we believed the diffusion coefficient

measurement technique is adsorption time dependent. Figure 23 shows the recorded

Page 78: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

63

pressure drop curve of a complete pressure drop when gas equilibrium pressure is ~4.10

MPa. Three data points marked in black “X” represent gas pressure recorded when the

adsorption process proceeded to 10, 20 and 30 minutes, respectively. The concluding

pressure for each point directly determines the adsorption amount at each recording time.

So by choosing different groups of pressure and adsorption time for adsorption amount and

diffusivity calculation in Equation (8), we may end up with different diffusion coefficient

results. The trends of the results are consistent with other diffusion results reported by.(C.

R. Clarkson & Bustin, 1999; Xiaojun Cui, Bustin, & Dipple, 2004; A. Kumar, 2007) The

values of diffusivity ranges from 7.22 × 10-13 to 9.18 × 10-15 m2/s. All diffusion coefficient

values kept decreasing with pressure drop from around 2 MPa to the highest pressure point

we reached in the experiments. During methane adsorption process, the pressure changing

rate reflects the significance of diffusion process. The decreasing rates are almost constant

and show linear relationship between gas pressure and diffusion coefficient. As we

discussed in adsorption results analysis part, the linkage between the significance of

adsorption and diffusion process is how adsorption, as a storage mechanism, controls the

gas molecule source and varies the gas concentration between cleats and micropores in

coal matrix. Figure 24 shows the gas molecule distribution around coal matrix and their

behavior when pressure changes. At low pressure in early stage of diffusion experiment,

the gas molecules outside coal matrix had higher concentration than those inside, creating

large concentration gradient in between. According to Fick’s law, a more significant gas

diffusion flow, shown as bold blue arrow lines, along the direction of entering the coal

matrix was observed. With gas pressure increases as experiment carrying on, the adsorption

site is received more gas molecules with increasing gas concentration. The high

Page 79: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

64

concentration of methane molecules inside coal matrix reduced the concentration gradient,

and thus hindered the diffusion flow, shown as red thin arrow lines, into the micropores.

The diffusion coefficient became extremely low at pressure larger than 8 MPa and can be

neglected if higher pressure is reached. The methane concentration inside coal matrix

became very high and the concentration gradient in between was reduced. Thus we can

safely expect a reduction in the significance of diffusion process when injection pressure

increases. On the other hand, from 0 to 2 MPa we also observed increase on all the diffusion

coefficient profiles. This phenomenon may be because at the very beginning of diffusion

measurement, the sample particles were vacuumed with no adsorbed gas molecules. On

the first step of gas injection, the transport of gas molecules may have faced with resistance

initially while in the next steps they were transporting in an expanded pore system. But

generally, negative correlations between diffusion coefficients and pressures are still

clearly observed. It explains that the non-Darcy components really take over a large

contribution on gas flow at low pressure.(Y. Wang, Liu, & Elsworth, 2015) In addition, by

comparing the adsorption data collected at different adsorption time points (10, 20 and 30

mins), we also observed and confirmed the influence of choosing different data recording

period on diffusion results. In Figure 21 and 22, although it’s hard to tell if there is a certain

correlation between recording time and diffusion coefficient, we can still capture the

differences among three chosen time. Theoretically, this calculation method we used

should yield same results at same pressure without specifying a certain time. However, the

final calculation result can be subjective over different researchers since they may not unify

this time parameter. In this measurement, we are able to at least have a baseline and idea

Page 80: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

65

about how the diffusion coefficient can potentially be varying due to adsorption recording

time.

4.3 Diffusive permeability results

According to the proposed model by Cui et al.(X. Cui et al., 2009) in Equation (3),

we obtained the equivalent permeability accounting for diffusion process in Figure 25 and

26. If we assume methane diffusion flow dominates, this calculated equivalent permeability

can be considered as gas permeability contributed only by diffusion process. If only

diffusion process is assumed in late-time CBM production, an increment in permeability

converted from diffusion coefficient is expected. Based on the result plots, we can observe

that there is also a negative correlation between equivalent permeability and gas pressure,

which is similar to the observation on the diffusion coefficient results. The permeability

values decreased with pressure increased, together with the increasingly intensified

diffusion process. For the equivalent permeabilities of two types of coal in different

adsorption recording time, when the pressure is higher than 6.5 MPa/940 psi, the

permeabilities are extremely small and can be ignored. When the pressure was lower than

2 MPa/280 psi, there was an obvious enhancement on every permeability profile. And this

is when the long production tail occurs and diffusion flow start to dominate.

For real CBM production process, gas permeability will eventually increase at low

pressure due to multiple rock deformation effect.(S. Liu & Harpalani, 2013b) When the

pore pressure keeps decreasing, a significant portion of sorbing gas will desorb from the

matrix and this sorption-induced permeability increase will dominate the flow behaviors at

low pressure.(Y. Wang et al., 2015) So generally gas permeability results are based on the

Page 81: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

66

combination of Darcy flow and non-Darcy flow including diffusion and rock deformation.

Compared to gas/apparent permeability results from some of the preceding works(Shugang

Wang et al., 2011; Y. Wang et al., 2015), the diffusion equivalent permeabilities in this

work have two main differences. First, the overall values of equivalent permeability are

quite low. Diffusion, in this case, seems to be less influential in coal gas mass flow

contribution. However, as we discussed previously in this work, even though Darcy

permeability can be high at low pressure, diffusion flow is the dominant factor that controls

the CBM production rate. The equivalent permeability values were only obtained based on

two assumptions that only diffusion flow occurs and Darcy permeability and diffusion

coefficient can be interchangeable. Plus, the gas rate at late-time production is usually quite

low, referring to Figure 16. The relatively low equivalent permeability we obtained

through calculation is able to give us a contrast to the situation when Darcy flow dominates.

Secondly, the equivalent permeabilities have increasing trends during CBM depletion,

which is determined by the intensity of diffusion process and gas flow properties. This

increasing trend in diffusive permeability only matches the increment appeared in sorbing

gas permeability results at low pressure in previous researchers’ works. This also explains

that it is non-Darcy process, such as adsorption and diffusion process, rather than pure

Darcy flow that has significant influence on low pressure permeability evolution(Y. Wang

et al., 2015). This information can be valuable in unconventional gas production prediction,

providing a different angle on how to interpret gas flow quality.

4.4 Coal type and diffusion

Figure 27 shows a comparison between diffusion coefficients at 30 minutes of two

different coal ranks from San Juan sub-bituminous coal and Pittsburgh bituminous coal

Page 82: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

67

samples. It can be easily observed that the San Juan sub-bituminous sample has higher

diffusion coefficient values, though sub-bituminous coal generally has lower carbon

content and energy efficiency. An et al.(An, Cheng, Wu, & Wang, 2013) revealed a

negative correlation between Langmuir pressure and coal rank. The increased micropore

volume with higher coal rank can accompany with lower Langmuir pressure, which

indicates a lower adsorption speed and smaller diffusion coefficient overall values. This

matches with the adsorption isotherm result in Figure 20 that San Juan coal sample has

higher adsorption capacity and, especially, higher adsorption rate at each pressure step. The

diffusive permeabilities between those two samples showed fairly similar difference in

magnitude in Figure 28. So in this study we may say that the significance of diffusion

process and diffusive permeability is mostly determined by the ad-/desorption behavior of

coal.

5. Conclusion

A series of experimental works and theoretical analysis on diffusion coefficient and

its equivalent permeability on San Juan sub-bituminous and Pittsburgh bituminous

pulverized coal samples has been presented. Unipore model and particle method was used

to measure and evaluate the diffusion coefficients with different pressure ranges and

different data collection time. An equivalent permeability was used to evaluate the quality

of gas transport assuming diffusion flow dominates. The results showed that diffusion

process is controlled by pressure and ad-/desorption process, and have large influence on

the quality of gas transport inside pore system of coal. Based on the work completed, the

following conclusions are made and summarized:

1. Diffusion coefficient results ranges from 7.22 x 10-13 to 9.18 x 10-15 m2/s obtained

Page 83: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

68

by using unipore model. It indicates a micropore scale diffusion behavior during

the tests.

2. The significance of diffusion flow depends on the ad-/desorption rate. At low

pressure (< 2MPa/280 psi) when the sorption process is severer, higher diffusion

coefficient values can be observed. Generally, a negative correlation between

diffusion coefficient and pressure is expected.

3. At each pressure step, more intensive diffusion flow can be captured in early time

adsorption rather than late time adsorption. Because the sorption rate kept

decreasing during the entire gas adsorption step.

4. The equivalent permeability converted from diffusion coefficient can reflect the

contribution of diffusion process in CBM gas transport if assumed diffusion process

dominates only. The overall value of diffusive permeability is quite low. At low

pressure (< 2MPa/280 psi), the diffusive permeability can match with preceding

experience on coal gas permeability since they all have increasing trends when

pressure keep decreasing.

5. San Juan sub-bituminous coal sample tends to have higher adsorption capacity and

sorption rate than Pittsburgh #8 bituminous coal sample during gas injection steps.

This leads to a more intensive diffusion process and higher quality of gas transport

in pores overall.

Page 84: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

69

References

Aminian, K., & Ameri, S. (2004). Type Curves for Coalbed Methane Production

Prediction. 2004 SPE Eastern Regional Meeting. http://doi.org/10.2523/91482-MS

Amyx, J. W., Bass, D. M. J., & Whiting, R. L. (1960). Petroleum Reservoir Engineering.

An, F. H., Cheng, Y. P., Wu, D. M., & Wang, L. (2013). The effect of small micropores

on methane adsorption of coals from Northern China. Adsorption, 19(1), 83–90.

http://doi.org/10.1007/s10450-012-9421-3

Brace, W. F., Walsh, J. B., & Frangos, W. T. (1968). Permeability of granite under high

pressure. Journal of Geophysical Research, 73(6).

Breckels, I. M., & Eekelen, H. a. M. (1982). Relationship Between Horizontal Stress and

Depth in Sedimentary Basins. Journal of Petroleum Technology, 34(9), 2191–2199.

http://doi.org/10.2118/10336-PA

Busch, A., Gensterblum, Y., Krooss, B. M., & Littke, R. (2004). Methane and carbon

dioxide adsorption–diffusion experiments on coal: upscaling and modeling.

International Journal of Coal Geology, 60(2–4), 151–168.

http://doi.org/10.1016/j.coal.2004.05.002

Cai, Y., Liu, D., Pan, Z., Yao, Y., Li, J., & Qiu, Y. (2013). Pore structure and its impact

on CH4 adsorption capacity and flow capability of bituminous and subbituminous

coals from Northeast China. Fuel, 103, 258–268.

http://doi.org/10.1016/j.fuel.2012.06.055

Choi, J.-G., Do, D. D., & Do, H. D. (2001). Surface Diffusion of Adsorbed Molecules in

Porous Media: Monolayer, Multilayer, and Capillary Condensation Regimes.

Industrial & Engineering Chemistry Research, 40(19), 4005–4031.

http://doi.org/10.1021/ie010195z

Civan, F., Rai, C., & Sondergeld, C. (2011). Shale-gas permeability and diffusivity

inferred by improved formulation of relevant retention and transport mechanisms.

Transport in Porous Media, 86(3), 925–944. http://doi.org/10.1007/s11242-010-

9665-x

Clarkson, C., & Bustin, R. (1999). 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, di, 1333–1344.

Clarkson, C., Pan, Z., Palmer, I., & Harpalani, S. (2010). Predicting sorption-induced

strain and permeability increase with depletion for coalbed-methane reservoirs. SPE

Journal, (March). R

Clarkson, C. R., & Bustin, R. M. (1999). The effect of pore structure and gas pressure

upon the transport properties of coal: a laboratory and modeling study. 2. Adsorption

rate modeling. Fuel, 78(11), 1345–1362. http://doi.org/10.1016/S0016-

2361(99)00056-3

Clarkson, C. R., Solano, N., & Bustin, R. (2013). Pore structure characterization of North

Page 85: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

70

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

intrusion. Fuel, 103, 606–616. http://doi.org/10.1016/j.fuel.2012.06.119

Coal Age - DEC 2014. (2014).

Coppens, M. (1999). The effect of fractal surface roughness on diffusion and reaction in

porous catalysts – from fundamentals to practical applications. Catalysis Today,

53(2), 225–243. http://doi.org/10.1016/S0920-5861(99)00118-2

Coppens, M. O., & Dammers, A. J. (2006). Effects of heterogeneity on diffusion in

nanopores - From inorganic materials to protein crystals and ion channels. Fluid

Phase Equilibria, 241(1–2), 308–316. http://doi.org/10.1016/j.fluid.2005.12.039

Crank, J. (1957). The mathematics of diffusion.

Cui, X., Bustin, A., & Bustin, R. (2009). Measurements of gas permeability and

diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 40479.

Cui, X., Bustin, A. M. M., & Bustin, R. M. (2009). Measurements of gas permeability

and diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 9(3), 208–223. http://doi.org/10.1111/j.1468-8123.2009.00244.x

Cui, X., & Bustin, R. M. (2005). Volumetric strain associated with methane desorption

and its impact on coalbed gas production from deep coal seams. AAPG Bulletin,

89(9), 1181–1202. http://doi.org/10.1306/05110504114

Cui, X., Bustin, R. M., & Dipple, G. (2004). Selective transport of CO2, CH4, and N2 in

coals: insights from modeling of experimental gas adsorption data. Fuel, 83(3), 293–

303. http://doi.org/10.1016/j.fuel.2003.09.001

Darabi, H., Ettehad, a., Javadpour, F., & Sepehrnoori, K. (2012). Gas flow in ultra-tight

shale strata. Journal of Fluid Mechanics, 710, 641–658.

http://doi.org/10.1017/jfm.2012.424

Dicker, A. I., & Smits, R. M. (1988). A practical approach for determining permeability

from laboratory pressure-pulse decay measurements. In SPE International Meeting

on Petroleum Engineering, Tianjin, China,1-4 Nov,1998. Tianjin, China.

Do, D. D. (1998). Adsorption Analysis: Equilibria and Kinetics (Vol. 2). Imperial

College Press. http://doi.org/10.1142/9781860943829

Drazer, G., Chertcoff, R., Bruno, L., & Rosen, M. (1999). Concentration dependence of

diffusion–adsorption rate in activated carbon. Chemical Engineering Science, 54,

4285–4291.

Dutta, A. (2009). Multicomponent Gas Diffusion and Adsorption in Coals for Enhanced

Methane Recovery.

EIA. (2015). Marcellus Region Drilling Productivity Report, (March), 2015.

Ertekin, T., King, G., & Schwerer, F. (1986). Dynamic Gas Slippage: A Unique Dual-

Mechanism Approach to the Flow of Gas in Tight Formations. SPE Formation

Evaluation. http://doi.org/10.2118/12045-PA

Page 86: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

71

Gan, H., Nandi, S. P., & Walker, P. L. (1972). Nature of the porosity in American coals.

Fuel, 51(4), 272–277. http://doi.org/10.1016/0016-2361(72)90003-8

Geertsma, J. (1966). Problems of rock mechanics in petroleum production engineering.

1st ISRM Congress, 3(59), 585–594.

Gensterblum, Y., Ghanizadeh, A., & Krooss, B. M. (2014). Gas permeability

measurements on Australian subbituminous coals: Fluid dynamic and poroelastic

aspects. Journal of Natural Gas Science and Engineering, 19, 202–214.

http://doi.org/10.1016/j.jngse.2014.04.016

Halliburton Company. (2007). Chapter 04 - Reservoir Analysis. Coalbed Methane

Princples and Practices, (June 2007), 191–282.

Harpalani, S., & Chen, G. (1997). Influence of gas production induced volumetric strain

on permeability of coal. Geotechnical and Geological Engineering, 15(4), 303–325.

http://doi.org/10.1007/BF00880711

Harpalani, S., & Mitra, A. (2009). Impact of CO2 Injection on Flow Behavior of Coalbed

Methane Reservoirs. Transport in Porous Media, 82(1), 141–156.

http://doi.org/10.1007/s11242-009-9475-1

Harpalani, S., Prusty, B., & Dutta, P. (2006). Methane/CO2 sorption modeling for

coalbed methane production and CO2 sequestration. Energy & Fuels, 20(4), 1591–

1599.

Hartman, R. (2008). Recent advances in the analytical methods used for shale gas

reservoir gas-in-place assessment. In AAPG Annual Convention, San Antonio,

Texas, 20-23 Apr, 2008 (Vol. 40317).

He, W., Lv, W., & Dickerson, J. (2014). Gas Diffusion Mechanisms and Models. In Gas

Transport in Solid Oxide Fuel Cells (pp. 9–17). http://doi.org/10.1007/978-3-319-

09737-4

Heller, R., Vermylen, J., & Zoback, M. (2014). Experimental investigation of matrix

permeability of gas shales. AAPG Bulletin, 98(5), 975–995.

http://doi.org/10.1306/09231313023

Hildenbrand, a., Krooss, B. M., Busch, a., & Gaschnitz, R. (2006). Evolution of

methane sorption capacity of coal seams as a function of burial history — a case

study from the Campine Basin, NE Belgium. International Journal of Coal Geology,

66(3), 179–203. http://doi.org/10.1016/j.coal.2005.07.006

Hsieh, P. a., Tracy, J. V., Neuzil, C. E., Bredehoeft, J. D., & Silliman, S. E. (1981). A

transient laboratory method for determining the hydraulic properties of “tight”

rocks—I. Theory. International Journal of Rock Mechanics and Mining Sciences &

Geomechanics Abstracts, 18(3), 245–252. http://doi.org/10.1016/0148-

9062(81)90979-7

Hubbert, M. K. (1957). Darcy’S Law and the Field Equations of the Flow of

Underground Fluids. International Association of Scientific Hydrology. Bulletin,

2(1), 23–59. http://doi.org/10.1080/02626665709493062

Page 87: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

72

Izadi, G., & Elsworth, D. (2013). The effects of thermal stress and fluid pressure on

induced seismicity during stimulation to production within fractured reservoirs.

Terra Nova, 25(5), 374–380. http://doi.org/10.1111/ter.12046

Izadi, G., Wang, S., Elsworth, D., Liu, J., Wu, Y., & Pone, D. (2011). Permeability

evolution of fluid-infiltrated coal containing discrete fractures. International Journal

of Coal Geology, 85(2), 202–211. http://doi.org/10.1016/j.coal.2010.10.006

Javadpour, F. (2009). Nanopores and apparent permeability of gas flow in mudrocks

(shales and siltstone). Journal of Canadian Petroleum Technology, 48, 1–6.

Javadpour, F., Fisher, D., & Unsworth, M. (2007). Nanoscale gas flow in shale gas

sediments. Journal of Canadian Petroleum Technology, 46(10).

http://doi.org/10.2118/07-10-06

Jones, S. (1997). A technique for faster pulse-decay permeability measurements in tight

rocks. SPE Formation Evaluation, (November 1994), 25–28.

Kamath, J., Boyer, R., & Nakagawa, F. (1992). Characterization of core-scale

heterogeneities using laboratory pressure transients. SPE Formation Evaluation,

(September), 219–227.

Karniadakis, G., Beskok, A., & Narayan, A. (2005). Microflows and Nanoflows.

http://doi.org/10.1007/0-387-28676-4

Klinkenberg, L. J. (1941). The permeability of porous media to liquids and gases.

Drilling and Production Practice, (2), 57–73.

http://doi.org/10.5510/OGP20120200114

Kuila, U., Prasad, M., & Kazemi, H. (2013). Assessing Knudsen flow in gas-flow models

of shale reservoirs (pp. 23–28).

Kumar, A. (2007). Methane Diffusion Characteristics of Illinois Coals.

Kumar, H., Elsworth, D., Liu, J., Pone, D., & Mathews, J. P. (2012). Optimizing

enhanced coalbed methane recovery for unhindered production and CO2 injectivity.

International Journal of Greenhouse Gas Control, 11, 86–97.

http://doi.org/10.1016/j.ijggc.2012.07.028

Kumar, H., Elsworth, D., Mathews, J. P., Liu, J., & Pone, D. (2014). Effect of CO2

injection on heterogeneously permeable coalbed reservoirs. Fuel, 135, 509–521.

http://doi.org/10.1016/j.fuel.2014.07.002

Kumar, H., Elsworth, D., Mathews, J. P., & Marone, C. (2015). Permeability evolution in

sorbing media: analogies between organic-rich shale and coal. Geofluids.

http://doi.org/10.1111/gfl.12135

Kwon, O., Kronenberg, A. K., Gangi, A. F., & Johnson, B. (2001). Permeability of

Wilcox shale and its effective pressure law. Journal of Geophysical Research, 106,

19339. http://doi.org/10.1029/2001JB000273

Langmuir, I. (1918). The adsorption of gases on plane surfaces of glass, mica and

platinum. Journal of the American Chemical Society, 345(1914).

Page 88: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

73

Levine, J. (1996). Model study of the influence of matrix shrinkage on absolute

permeability of coal bed reservoirs. Geological Society, London, Special

Publications.

Li, J., Liu, D., Lu, S., Yao, Y., & Xue, H. (2015). Evaluation and modeling of the CO2

permeability variation by coupling effective pore size evolution in anthracite coal.

Energy & Fuels. http://doi.org/10.1021/ef502674y

Li, J., Liu, D., Yao, Y., Cai, Y., & Chen, Y. (2013). Evaluation and modeling of gas

permeability changes in anthracite coals. Fuel, 111, 606–612.

http://doi.org/10.1016/j.fuel.2013.03.063

Li, Y., Li, X., Shi, J., Wang, H., Wu, L., Teng, S., & Corp, S. (2014). A Nano-Pore Scale

Gas Flow Model for Shale Gas Reservoir Flow characteristic of shale gas reservoir

in various pole scales and molecule scales, (June), 9–11.

Liu, J., Chen, Z., Elsworth, D., Miao, X., & Mao, X. (2010). Linking gas-sorption

induced changes in coal permeability to directional strains through a modulus

reduction ratio. International Journal of Coal Geology, 83(1), 21–30.

http://doi.org/10.1016/j.coal.2010.04.006

Liu, S. (2012). Estimation of different coal compressibilities of coalbed methane

reservoirs under replicated in situ condition.

Liu, S., & Harpalani, S. (2013a). A new theoretical approach to model sorption-induced

coal shrinkage or swelling. AAPG Bulletin, 97(7), 1033–1049.

http://doi.org/10.1306/12181212061

Liu, S., & Harpalani, S. (2013b). Permeability prediction of coalbed methane reservoirs

during primary depletion. International Journal of Coal Geology, 113, 1–10.

http://doi.org/10.1016/j.coal.2013.03.010

Liu, S., & Harpalani, S. (2014a). Compressibility of sorptive porous media: Part 1.

Background and theory. AAPG Bulletin, c, 1–25.

Liu, S., & Harpalani, S. (2014b). Compressibility of Sorptive Porous Media: Part I –

Background and Theory. AAPG Bulletin, 98(9), 1761–1772.

Liu, S., & Harpalani, S. (2014c). Compressibility of Sorptive Porous Media: Part II –

Experimental Study on Coal. AAPG Bulletin, 98(9), 1773–1788.

Liu, S., & Harpalani, S. (2014d). Determination of the effective stress law for

deformation in coalbed methane reservoirs. Rock Mechanics and Rock Engineering,

47(5), 1809–1820. http://doi.org/10.1007/s00603-013-0492-6

Liu, S., & Harpalani, S. (2014e). Evaluation of in situ stress changes with gas depletion

of coalbed methane reservoirs. Journal of Geophysical Research: Solid Earth,

119(8), 6263–6276. http://doi.org/10.1002/2014JB011228.

Liu, S., Harpalani, S., & Pillalamarry, M. (2012). Laboratory measurement and modeling

of coal permeability with continued methane production: Part 2 – Modeling results.

Fuel, 94, 117–124. http://doi.org/10.1016/j.fuel.2011.10.053

Lorenz, J. C., Teufel, L. W., & Warpinski, N. R. (1991). Regional fractures; I, A

Page 89: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

74

mechanism for the formation of regional fractures at depth in flat-lying reservoirs.

AAPG Bulletin , 75(11), 1714–1737.

Luffel, D., Hopkins, C., & Jr, P. S. (1993). Matrix permeability measurement of gas

productive shales. In SPE Annual Technical Conference and Exhibition, Houston,

Texas, 2-6 Oct, 1993.

Ma, Q., Harpalani, S., & Liu, S. (2011). A simplified permeability model for coalbed

methane reservoirs based on matchstick strain and constant volume theory.

International Journal of Coal Geology, 85(1), 43–48.

http://doi.org/10.1016/j.coal.2010.09.007

Malkovsky, V. I., Zharikov, a. V., & Shmonov, V. M. (2009). New methods for

measuring the permeability of rock samples for a single-phase fluid. Izvestiya,

Physics of the Solid Earth, 45(2), 89–100.

http://doi.org/10.1134/S1069351309020013

Markowski, A. K. (2014). Pennsylvania Coalbed Methane Update, 80353.

Markowski, A., Resources, N., & Virginia, W. (2014). Coalbed methane in

Pennsylvania.

Mavor, M. J., Owen, L. B., & Pratt, T. J. (1990). Measurement and Evaluation of Coal

Sorption Isotherm Data. Proceedings of SPE Annual Technical Conference and

Exhibition, (1). http://doi.org/10.2523/20728-MS

Mazumder, S., & Wolf, K. (2008). Differential swelling and permeability change of coal

in response to CO 2 injection for ECBM. International Journal of Coal Geology.

Mckernan, R. E., Rutter, E. H., Mecklenburgh, J., & Taylor, K. G. (2014). Influence of

Effective Pressure on Mudstone Matrix Permeability: Implications for Shale Gas

Production. In SPE/EAGE European Unconventional Conference and Exhibition,

Vienna, Austria, 25-27 Feb, 2014 (p. 13). Society of Petroleum Engineers.

http://doi.org/10.2118/167762-MS

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1–Laboratory results. Fuel,

94, 110–116.

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1 – Laboratory results. Fuel,

94, 110–116. http://doi.org/10.1016/j.fuel.2011.10.052

Okotie, V. U., & Moore, R. L. (2011). Well-production challenges and solutions in a

mature, very-low-pressure coalbed-methane reservoir. SPE Production and

Operations, 26(2), 149–161. http://doi.org/10.2118/107705-MS

Palmer, I., & Mansoori, J. (1998). How permeability depends on stress and pore pressure

in coalbeds: a new model. SPE Reservoir Evaluation and Engineering, SPE

52607(December), 539–544.

Palmer, S. I., & Mansoori, J. (1998). How permeability depends on stress and pore

pressure in coalbeds: a new model. Soc Petrol Eng Res Eng, (December), 539–544.

Page 90: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

75

Pan, Z., & Connell, L. D. (2007). A theoretical model for gas adsorption-induced coal

swelling. International Journal of Coal Geology, 69(4), 243–252.

http://doi.org/10.1016/j.coal.2006.04.006

Pillalamarry, M., Harpalani, S., & Liu, S. (2011). Gas diffusion behavior of coal and its

impact on production from coalbed methane reservoirs. International Journal of

Coal Geology, 86(4), 342–348. http://doi.org/10.1016/j.coal.2011.03.007

Randolph, P. L., Soeder.D.J., & Chowdiah, P. (1984). Porosity and Permeability of Tight

Sands. In SPE/DOE/GRI Unconventional Gas Recovery Symposium held in

Pittsburgh, PA, May 13-15. Dallas.

Reif.F. (1965). Fundamentals of statistical and thermal physics.

Rexer, T. F. T., Benham, M. J., Aplin, A. C., & Thomas, K. M. (2013). Methane

Adsorption on Shale under Simulated Geological Temperature and Pressure

Conditions. Energy & Fuels, 27(6), 3099–3109. http://doi.org/10.1021/ef400381v

Rodrigues, C. F., & Lemos De Sousa, M. J. (2002). The measurement of coal porosity

with different gases. International Journal of Coal Geology, 48(3–4), 245–251.

http://doi.org/10.1016/S0166-5162(01)00061-1

Roy, S., Raju, R., Chuang, H. F., Cruden, B. a., & Meyyappan, M. (2003). Modeling gas

flow through microchannels and nanopores. Journal of Applied Physics, 93(8),

4870–4879. http://doi.org/10.1063/1.1559936

Ruckenstein, E. (1971). Sorption by solids with bidisperse pore structures. Chemical

Engineering …, 26, 1305–1318.

Rydzy, M. B., Patino, J., Elmetni, N., & Appel, M. (2016). Stressed Permeability in

Shales : Effects of Matrix Compressibility and Fractures - A Step towards

Measuring Matrix Permeability in Fractured Shale Samples, 1–12.

Sakhaee-Pour, A., & Bryant, S. (2012). Gas Permeability of Shale. SPE Reservoir

Evaluation & Engineering, 15, 401–409.

Seidle, J. P., Jeansonne, M. W., & Erickson, D. J. (1992). Application of matchstick

geometry to stress dependent permeability in coals. SPE Rocky Mountain Regional

Meeting, 433–444. http://doi.org/10.2523/24361-MS

Shabro, V. (2012). Forecasting Gas Production in Organic Shale with the Combined

Numerical Simulation of Gas Diffusion in Kerogen, Langmuir Desorption from

Kerogen Surfaces, and Advection in Nanopores. SPE Annual Technical ….

http://doi.org/10.2118/159250-MS

Shi, J.-Q., Pan, Z., & Durucan, S. (2014). Analytical models for coal permeability

changes during coalbed methane recovery: Model comparison and performance

evaluation. International Journal of Coal Geology, 136, 17–24.

http://doi.org/10.1016/j.coal.2014.10.004

Shi, J., & Durucan, S. (2005). A model for changes in coalbed permeability during

primary and enhanced methane recovery. SPE Reservoir Evaluation & Engineering,

8(4), 291–299.

Page 91: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

76

Shi, J. Q., & Durucan, S. (2003). A bidisperse pore diffusion model for methane

displacement desorption in coal by CO 2 injection☆. Fuel, 82(10), 1219–1229.

http://doi.org/10.1016/S0016-2361(03)00010-3

Shi, J. Q., & Durucan, S. (2004). Drawdown Induced Changes in Permeability of

Coalbeds: A New Interpretation of the Reservoir Response to Primary Recovery.

Transport in Porous Media, 56(1), 1–16.

http://doi.org/10.1023/B:TIPM.0000018398.19928.5a

Shi, J. Q., & Durucan, S. (2014). Modelling laboratory horizontal stress and coal

permeability data using S&D permeability model. International Journal of Coal

Geology, 131, 172–176. http://doi.org/10.1016/j.coal.2014.06.014

Singh, H., & Javadpour, F. (2013). Nonempirical Apparent Permeability of Shale. In

Unconventional Resources Technology Conference (Vol. 2, p. 16). Denvor,

Colorado.

Singh, H., & Javadpour, F. (2015). Langmuir slip-Langmuir sorption permeability model

of shale. FUEL, (October). http://doi.org/10.1016/j.fuel.2015.09.073

Smith, D., & Williams, F. (1984). Diffusional effects in the recovery of methane from

coalbeds. SPEJ, Soc. Pet. Eng. J.;(United States), (October), 529–535.

Soeder, D. (1988). Porosity and permeability of eastern Devonian gas shale. SPE

Formation Evaluation, 3, 116–124.

Sudibandriyo, M., Pan, Z., Fitzgerald, J. E., Jr, R. L. R., & Gasem, K. A. M. (2003).

Adsorption of Methane, Nitrogen, Carbon Dioxide, and Their Binary Mixtures on

Dry Activated Carbon at 318.2 K and Pressures up to 13.6 MPa. Lan, 19(8), 5323–

5331.

Suggested Methods for Uniaxial‐Strain Compressibility Testing for Reservoir

Geomechanics. (2015), c(April), 1–29.

Swami, V. (2012). Shale Gas Reservoir Modeling: From Nanopores to Laboratory. SPE

Annual Technical Conference and Exhibition, (October), 8–10.

Swami, V., Clarkson, C., & Settari, A. (2012). Non-Darcy Flow in Shale Nanopores: Do

We Have a Final Answer? SPE Canadian Unconventional …, (November), 1–17.

Swami, V., & Settari, A. (2012). A pore scale gas flow model for shale gas reservoir. In

SPE Americas Unconventional Resources Conference held in Pittsburgh (pp. 5–7).

Terzaghi, K. (1943). Theoretical Soil Mechanics. John Wiley & Sons, Inc.

http://doi.org/10.1680/geot.1964.14.1.1

Teufel, L. W., Rhett, D. W., & Farrell, H. E. (1991). Effect of Reservoir Depletion And

Pore Pressure Drawdown On In Situ Stress And Deformation In the Ekofisk Field,

North Sea. The 32nd U.S. Symposium on Rock Mechanics (USRMS), 63–72.

U S Energy Information Administration. (2011). Emissions of Greenhouse Gases in the

United States 2009, (March).

Walls, J., Diaz, E., & Cavanaugh, T. (2012). Shale Reservoir Properties from Digital

Page 92: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

77

Rock Physics. Proceedings of SPE/EAGE European Unconventional Resources

Conference and Exhibition, 1–9. http://doi.org/10.2118/152752-MS

Wang, G., Ren, T., Wang, K., & Zhou, A. (2014). Improved apparent permeability

models of gas flow in coal with Klinkenberg effect. Fuel, 128, 53–61.

http://doi.org/10.1016/j.fuel.2014.02.066

Wang, S., Elsworth, D., & Liu, J. (2011). Permeability evolution in fractured coal: the

roles of fracture geometry and water-content. International Journal of Coal

Geology, 87(1), 13–25. http://doi.org/10.1016/j.coal.2011.04.009

Wang, S., Elsworth, D., & Liu, J. (2012). A mechanistic model for permeability evolution

in fractured sorbing media. Journal of Geophysical Research, 117(B6).

Wang, S., Elsworth, D., & Liu, J. (2012a). A mechanistic model for permeability

evolution in fractured sorbing media. Journal of Geophysical Research: Solid Earth

( …, 117(B6), B06205. http://doi.org/10.1029/2011JB008855

Wang, S., Elsworth, D., & Liu, J. (2012b). Permeability evolution during progressive

deformation of intact coal and implications for instability in underground coal

seams. International Journal of Rock Mechanics and Mining Sciences, 1–12.

Wang, Y., & Liu, S. (2016). Estimation of Pressure-Dependent Diffusive Permeability of

Coal Using Methane Diffusion Coefficient: Laboratory Measurements and

Modeling. Energy and Fuels, 30(11), 8968–8976.

http://doi.org/10.1021/acs.energyfuels.6b01480

Wang, Y., Liu, S., & Elsworth, D. (2015). Laboratory investigations of gas flow

behaviors in tight anthracite and evaluation of different pulse-decay methods on

permeability estimation. International Journal of Coal Geology, 149, 118–128.

http://doi.org/10.1016/j.coal.2015.07.009

Wu, K., Li, X., Wang, C., Chen, Z., & Yu, W. (2015). A Model for Gas Transport in

Microfractures of Shale and Tight Gas Reservoirs. AIChE Journal, 61(6).

http://doi.org/10.1002/aic

Wu, K., Li, X., Wang, C., Yu, W., & Chen, Z. (2014). Apparent Permeability for Gas

Flow in Shale Reservoirs Coupling Effects of Gas Diffusion and Desorption.

Proceedings of the 2nd Unconventional Resources Technology Conference, 1–18.

http://doi.org/10.15530/urtec-2014-1921039

Xu, H., Tang, D., Zhao, J., Li, S., & Tao, S. (2015). A new laboratory method for

accurate measurement of the methane diffusion coefficient and its influencing

factors in the coal matrix. Fuel, 158, 239–247.

http://doi.org/10.1016/j.fuel.2015.05.046

Yi, J., Akkutlu, I., & Deutsch, C. (2008). Gas Transport in Bidisperse Coal Particles:

Investigation for an Effective Diffusion Coefficient in Coalbeds. Journal of

Canadian Petroleum Technology, 47(10). http://doi.org/10.2118/08-10-20

Yin, G., Jiang, C., Wang, J. G., & Xu, J. (2013). Combined Effect of Stress, Pore

Pressure and Temperature on Methane Permeability in Anthracite Coal: An

Page 93: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

78

Experimental Study. Transport in Porous Media, 100(1), 1–16.

http://doi.org/10.1007/s11242-013-0202-6

Yuan, W., Pan, Z., Li, X., Yang, Y., Zhao, C., Connell, L. D., … He, J. (2014).

Experimental study and modelling of methane adsorption and diffusion in shale.

Fuel, 117, 509–519. http://doi.org/10.1016/j.fuel.2013.09.046

Zhang, T., Ellis, G. S., Ruppel, S. C., Milliken, K., & Yang, R. (2012). Effect of organic-

matter type and thermal maturity on methane adsorption in shale-gas systems.

Organic Geochemistry, 47, 120–131.

http://doi.org/10.1016/j.orggeochem.2012.03.012

Zoback, M. D., & Byerlee, J. D. (1975). Permeability and Effective Stress. Aapg Bulletin-

American Association of Petroleum Geologists, 59(1), 154–158.

http://doi.org/10.1306/83D91C40-16C7-11D7-8645000102C1865D

Appendix

Figure 16 Typical CBM production curve with relative volumes of methane and water through

production time.

Page 94: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

79

Figure 17 Gas desorption, diffusion and Darcy viscous flow in coal matrix.

Figure 18 Picture of the experimental setup for adsorption and diffusion measurement. Pulverized

coal samples are stored in the sample cell.

Page 95: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

80

Figure 19 A schematic view of sorption/diffusion experimental setup

Figure 20 CH4 absolute adsorption isotherms and Langmuir modeling result of San Juan sub-

Page 96: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

81

bituminous and Pittsburgh #8 bituminous pulverized dry samples.

Figure 21 San Juan coal sample CH4 diffusion coefficient variation with pressure.

Figure 22 Pittsburgh #8 coal sample CH4 diffusion coefficient variation with pressure.

Page 97: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

82

Figure 24 Comparison of gas molecule behavior in low and high pressure diffusion.

4.1

4.11

4.12

4.13

4.14

4.15

4.16

4.17

4.18

4.19

4.2

0 100 200 300 400 500 600 700 800 900 1000

Pre

ssu

re (

MP

a)

Time (min)

4.14

4.15

4.16

4.17

4.18

4.19

4.2

0 10 20 30 40 50 60 70 80 90 100

Pre

ssu

re (

MP

a)

Time (min)

High Gas

PressureLow Gas

Pressure

Figure 23 Pressure drop curve during diffusion coefficient measurement with initial pressure at 4.19

MPa. The 10, 20 and 30 minutes recording point are shown as black “X”.

Page 98: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

83

Figure 25 San Juan coal sample CH4 permeability evolution correlated to diffusion coefficient.

0

2

4

6

8

10

12

0 2 4 6 8 10

Dif

fusi

ve

Per

mea

bil

ity

(m

2)

Pressure (MPa)

10 min

20 min

30 min

×𝟏𝟎−26

Page 99: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

84

Figure 26 Pittsburgh #8 coal sample CH4 permeability evolution correlated to diffusion coefficient.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 2 4 6 8 10

Dif

fusi

ve

Per

mea

bil

ity

(m

2)

Pressure (MPa)

10 min

20 min

30 min

×𝟏𝟎−26

Page 100: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

85

Figure 27 Diffusion coefficient data comparison between San Juan and Pittsburgh #8 coal samples

in average values.

0

1

2

3

4

5

6

7

0 2 4 6 8 10

Dif

fusi

on

Co

effi

cien

t (m

2/s

)

Pore Pressure (MPa)

San Juan

Pittsburgh #8

×𝟏𝟎−13

Page 101: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

86

Figure 28 Comparison of CH4 permeability evolution correlated to diffusion coefficient between

San Juan and Pittsburgh #8 coal sample.

Table 5 Proximate analysis of San Juan sub-bituminous and Pittsburgh #8 bituminous coal samples.

Proximate Analysis

Sample Fixed carbon Moisture Ash content Volatile matter

San Juan 40.75% 4.66% 19.17% 35.42%

Pittsburgh#8 55.02% 2.13% 3.84% 39.01%

0

1

2

3

4

5

6

7

8

0 2 4 6 8 10

Eq

uiv

ale

nt

Per

mea

bil

ity

(m

2)

Pressure (MPa)

San Juan

Pittsburgh #8

×𝟏𝟎−26

Page 102: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

87

Chapter 4 Anisotropy Characteristics of Coal Shrinkage/Swelling and Its

Impact on Coal Permeability Evolution with CO2 Injection

Abstract

Coal deforms with gas depletion during coalbed methane (CBM) primary recovery and

CO2 storage for purposes of sequestration or enhanced recovery. The sorption-induced

deformation associated with gas adsorption and desorption has been experimentally and

theoretically studied and is considered to be a unique feature of coal. The results of a series

of experimental measurements of coal deformation with gas injection and depletion using

helium, methane, CO2, and various mixtures of methane and CO2 revealed that the coal

sorption induced deformation exhibits anisotropy, with larger deformation in direction

perpendicular to bedding than those parallel to the bedding planes. Furthermore, the

deformation of coal is reversible for helium and methane with injection/depletion, but not

for CO2. Aiming at improving the understanding of enhanced CBM process, a coal

deformation experimental study on methane displacement with CO2 was performed as

well, where incremental swelling was measured for each CO2 displacement step. The total

induced swelling strain (1.68%) with CO2 saturation was less than that for pure CO2

injection induced strain (1.87%) at same pressure of 5.9 MPa (850 psi). The measured

deformation with CO2 injection was applied to an analytical coal permeability model, of

Ma et al., which is capable of accommodating the anisotropic sorption-induced

deformation. Based on the modeling results, it was found that application of isotropic

deformation in permeability model can overestimate the permeability loss compared to

anisotropic deformation. This demonstrates that the anisotropic coal deformation should

Page 103: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

88

be considered to predict the permeability behavior of CBM as well as CO2

sequestration/ECBM projects.

1. Introduction

Coalbed methane (CBM) has become an important part of the world’s natural gas

resource. The energy industry classifies coal as an “unconventional gas reservoir” and has

worked continuously towards technologies to economically extract gas production from it.

Most recent data from the Energy Information Administration (EIA) state that proven CBM

reserves of the US accounted for about 17.5 trillion cubic feet (TCF) in 2010. CBM

production in 2010 was 1.9 TCF, which was approximately 8.5% of US natural gas

production 1 although this has slid recently as a result of a move favoring shale gas

production. In addition to the primary CBM production, carbon sequestration in deep and

unmineable coal is considered a promising technology that is feasible in the immediate

future for storage of carbon dioxide. Coupled with the potential of enhancing methane

recovery from gas producing coalbeds, it offers an economic advantage over all other

methods, e.g., deep saline aquifers and depleted natural gas reservoirs. This unique process

is known as enhanced coalbed methane recovery (ECBM). Several ECBM pilots have been

conducted around the world, including the Allison Unit in San Juan basin in the US 2,

MOVECBM project (previously RECOPOL in Poland 3, Yubari site in Japan 4, Fenn Big

Valley site in Canada 5 and Qinshui basin in China 6. Most recently, another ECBM pilot

test was carried out at the Pump Canyon site in San Juan basin with specific goal of

improving the understanding of the efficacy of enhanced coalbed methane recovery

processes via CO2 injection into a pressure depleted coal seam 7,8. The above pilot tests

Page 104: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

89

have shown that there is a significant economic and environmental benefit associated with

CO2-ECBM. However, several challenges remain and these must be overcome before the

technology is widely accepted by the industry. An incomplete understanding of the flow

behavior under in situ stress-strain conditions, concerns over future coal mineability,

permanence of storage, and formation stability are the major barriers towards acceptance

of coal as permanent and possible CO2 repositories.

The flow behavior in coal has been extensively investigated during the past two

decades. The fracture/cleat permeability evolution of coal depends on the net effect of

sorption-induced deformation and effective stress during both primary depletion and

ECBM. For primary methane recovery, the matrix shrinkage strain associated with gas

desorption tends to open the cleat aperture and thus increase the fracture permeability with

continued depletion 9-13. Conversely, CO2 injection induced strain that results in

permeability decrease due to partial closure of the cleats 10,14,15. The permeability of coal

is also influenced by the reservoir boundary conditions. Two boundaries were assumed for

different analytical coal permeability models, namely, uniaxial strain and constant volume

conditions, respectively 12. In this study, experimental work was carried out to measure the

matrix deformations for non-absorbing gas, helium, and sorbing gases, methane and CO2,

and their mixture. Following the experimental work, the influence of anisotropic sorption

induced deformation on permeability evolution of coal during CO2 injection under constant

volume theory was analyzed.

1.1 Deformation of coal matrix and fracture to pressure/stress

It has long been documented that gas pressure variation induces volumetric strain in

coal matrix 16-20 and thus changes the flow behavior of coal under stress-strain controlled

Page 105: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

90

conditions 9,12-15,20-23. Coal, known as a dual-porosity medium, consists of micro- and

macro- pores. The micro-pores serve as the storage house for the gas, where the amount of

adsorbed gas is typically quantified by the Langmuir model 24. The macro-pore system,

also termed as coal cleat system, provides the flow paths during gas depletion and injection.

Associated with gas ad-/de- sorption, the swelling or shrinkage of solid coal matrix is an

interesting and unique phenomenon, with a significant influence on the coal cleat system

and hence, coal permeability. The permeability of coal, to some extent, not only determines

the overall recovery performance during primary development but also has important effect

on CO2 injectivity during ECBM operations. The interaction between coal matrix and cleat

dominates the permeability evolution of coal under in situ condition. In the following

subsections, a brief discussion about the nature of some of these factors and evidence that

some of these are responsible for permeability variation, is provided.

1.2 Coal cleat network geometry

Cleats are natural fractures in coal and are formed during the process of coalification.

These account for most of the permeability of CBM reservoirs and can have a significant

impact on the success of engineering procedures, such as, cavity stimulation and reservoir

simulation. Given that the cleat system is irregular and complex, it is impossible to

precisely describe the cleat geometry in detail. However, the cleats are generally oriented

orthogonally 25. For the purpose of modeling, the conceptual models are essential in order

to quantify the changes in the fracture network and, hence, the permeability. Warren and

Root 26 introduced a conceptual cubic-geometry model for the dual porosity matrix-fracture

reservoir rocks, shown in Figure 29(a). This model was adopted in CBM operations in the

eighties 27,28. A more simplified and realistic model, bundle of matchstick geometry,

Page 106: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

91

illustrated in Figure 29(b), was first introduced by Seidle et al. 29 and then widely accepted

and used by others 9,10,12,14,30. Recently, both cubic and matchstick geometry models have

been criticized due to the oversimplification of the solid-fracture system 21,22,31 and a more

realistic physical structure was proposed as one where the coal matrix blocks are linked

with each other through “solid bridge(s)”. It is unquestionably more representative than the

earlier models although its application to theoretical modeling of permeability has technical

challenges, requiring a few parameters that are difficult to estimate and quantification. In

order to eliminate the use of fuzzy parameters involved in the “bridge model”, the bundle

of matchsticks geometry is used in this study to analyze the influence of coal deformation

on the evolution of permeability.

1.3 Anisotropy of matrix deformation

Directional deformation of coal matrix has been observed in several laboratory

studies15,19,32,33. In 1996, Levine reported that sorption-induced linear strain was slightly

higher perpendicular to bedding planes than in parallel direction for both methane and CO2

for Illinois high volatile C bituminous coal. Based on data reported by Levine, the linear

strain in direction perpendicular to bedding is five percent larger than that in the parallel

direction, which was relatively small compared to data reported in later studies 15,19,33. Day

et al. 19 used an optical technique to measure the strain of coal at different gas pressures

and reported that expansion in the direction perpendicular to bedding was 30-70% greater

than in parallel direction when three Australian bituminous coal samples were saturated

with CO2. Majewska et al. 33 presented a set of experimental results showing similar

anisotropic swelling phenomenon where strain perpendicular to the bedding was larger

than that in the parallel direction. Following this, Pan and Connell 15 reported that swelling

Page 107: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

92

strain in the direction parallel to bedding is between 52% and 60% of that in the direction

perpendicular to the bedding for methane and CO2 up to pressures of ~11 MPa. The

experimental data was successfully modeled using the Pan and Connell swelling approach

18, which is based on three-tube structure of coal and energy balance theory. Regardless of

the magnitude of the degree of anisotropy on coal swelling for different coal types, a

general conclusion can be made that the swelling effect in the direction perpendicular to

bedding is stronger than that parallel to bedding direction. The sorption-induced strain, to

some extent, controls the cleat aperture and directly impacts the permeability of coal. The

anisotropic behavior, therefore, obviously warrants more attention for both primary CBM

and ECBM operations.

1.4 Gas mixture sorption

Coal typically contains multicomponent gas mixtures, including methane, CO2, N2 and

other heavy hydrocarbons. Strictly speaking, the primary development of CBM involves

multicomponent gas desorption and it has been observed that CO2 concentration in the

produced gas increases with time over the life of wells in San Juan basin 7. However, most

of previous studies related to primary development only focus on single gas, methane,

desorption behavior and its impact on shrinkage and swelling. The potential of carbon

sequestration in subsurface coal has boosted the research in the area of CO2 adsorption on

coal and its ramifications, including coal swelling, permeability variations, coal softening

with CO2 adsorption, etc. CO2 has a higher affinity than methane for different coal types,

as reported by several researchers 2,15,34-38. As for ECBM, several investigations of gas

mixture induced deformation have been carried out. In 2008, Zarebska and Ceglarska-

Stefanska measured volumetric strain in Polish bituminous coal when exposed to methane,

Page 108: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

93

CO2 and mixture of the two gases. For pure gas, the swelling of coal when exposed to CO2

was approximately twice that for methane. For gas mixture containing these two gases, the

swelling is substantially lower than that for pure CO2 and the linear strain is slightly larger

than that for pure methane, even for 75% CO2 concentration 39. Following this, Harpalani

and Mitra 40 carried out an experimental study to measure the incremental strain with

methane displacement by CO2 injection. A significant incremental strain was observed for

both Illinois and San Juan coals. However, after completion of CO2-displacement, the

volumetric strain was slightly smaller than that induced by pure CO2, which suggests that

either some residual methane cannot be replaced by CO2, or that the physical structure of

coal changed as a result of gas adsorption. Unfortunately, no directional swelling strain

was reported. Recently, Day et al. 41 reported swelling of coal using a mixture of methane

and CO2. The swelling induced strain by the mixture of fixed composition was between

that of pure CO2 and methane, and was proportional to the concentration of CO2 in the

mixture. Additionally, the swelling of coal was primarily dominated by the partial

pressures of the component gases.

2. Laboratory experiments

2.1 Sample preparation

Coal deformation tests were carried out using coal samples from the Illinois basin.

Blocks of coal were collected from an underground coal mining operation. The coal blocks

were kept immersed in water to prevent weathering and oxidation. One two-inch (~5 cm)

cylindrical core was drilled from the coal blocks. The top and bottom part of core was

trimmed off and the middle portion was preserved for other purposes. The upper part of

the core was split into four quadrants. The height of each quadrant is ~ 4cm and the radius

Page 109: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

94

of each quadrant is ~ 2.5cm as shown in Figure 2. Three quadrants with smooth surfaces

perpendicular to each other were selected. Subsequent to preparation, the specimens were

kept in an environmental chamber under controlled conditions of temperature and humidity

until initiating the experiment.

2.2 Experimental setup

The experimental setup using strain gauges was designed and employed to measure

the sorption-induced strain in the three directions. This has been described in detail in a

previous publication 20. To enable achieving methane displacement by CO2, the original

setup was modified to include a gas chromatograph (GC) and a means to collect gas

samples from the high pressure vessels. The testing rig included four identical high

pressure vessels as discussed in a previous publication 20. Three were used in this study

and one of these three was installed with a trace gas collection system connecting it to the

GC for gas composition analysis. Figure 30 shows the schematic of the experimental setup,

along with the GC assembly. The main components of the setup were high pressure vessel,

data acquisition system, trace gas collection system and GC. All three high pressure vessels

were placed in a constant temperature water bath to eliminate any thermal impact on the

deformation of coal.

2.3 Experimental procedure

The three quadrants of coal were taken out of the environmental chamber. The testing

commenced with strain gauge affixation. For each sample, three strain gauges were

installed to the surfaces of each sample in order to monitor the linear strains in the three

orthogonal directions. Gauge #1 was perpendicular to the bedding plane and the other two

(#2 and #3) were parallel to bedding, as shown in Figure 30. The gauges were affixed

Page 110: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

95

using an epoxy recommended by the manufacturer. After attaching the gauges, lead wires

were soldered to each strain gauge. The entire assembly was then placed in the pressure

vessel with the required outlets from gauge connection to a data acquisition system, as

illustrated in Figure 30. This gauge installation procedure was repeated for all three

samples. After achieving temperature equilibrium, all three samples were first subjected to

increasing helium pressure in steps of ~1.4 MPa (200 psi) to a final pressure of ~6.9 MPa

(1000 psi). Helium being non-adsorptive, the deformation was purely due to mechanical

compression of the solid coal resulting from changes in the external stress/pressure. The

helium was then bled out in a step-wise manner down to near zero. During the helium

cycle, the linear strains of coal were continuously monitored and recorded for further

analysis.

After completing the helium cycle, one sample (#1) was subjected to step-wise

flooding with CO2 by increasing the pressure in steps to 1 MPa (150 psi), 2.1 MPa (300

psi), 3.1 MPa (450 psi), 4.5 MPa (650 psi) and a final pressure of 5.9 MPa (850 psi).

Following this, CO2 pressure was decreased down to near zero in a step-wise manner. The

other two samples were initially injected with methane at ~1 MPa (150 psi) and

subsequently at 2.1 MPa (300 psi), 3.4 MPa (500 psi), 4.8 MPa (700 psi), and finally, 6.9

MPa (1000 psi). At each pressure step, the coal samples were allowed to equilibrate. Of

these two, one (#3) was connected to the trace gas collection system and GC. Sample #2

was simply depressurized to near zero in a step-wise manner. The pressure vessel with

connection to GC was used to measure the incremental strain due to CO2 displacement.

Since the maximum pressure of CO2 cylinder is 6.2 MPa (900 psi), a constant pressure of

5.9 MPa (850 psi) was determined to be adequate for displacing methane with CO2. The

Page 111: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

96

methane pressure, therefore, was first decreased to 5.9 MPa (850 psi) and allowed to

equilibrate. The sample was then subjected to increasing concentration of CO2, maintaining

the total pressure constant at 5.9 MPa (850 psi). This was achieved by injecting CO2 while

bleeding methane/(methane + CO2) mixture out every time CO2 was injected. This ensured

that the solid and cleats in the samples were subjected to the same external conditions

throughout the test, the only difference being the composition of the gases within and

around the sample. The procedure was continued until the gas in the vessel was nearly pure

CO2. Hence, the sample was subjected to varying composition of CO2, all the way from 0

to 100%. At each equilibrium step, a sample of gas was analyzed using the GC and

concentration of CO2 was measured. The linear strains of each sample were continuously

recorded and subsequently used for analysis.

3. Experimental results and discussion

3.1 Helium injection results

The first part of this experiment involved dosing all three samples with increasing

amount of helium in pressure steps to a maximum of 6.9 MPa (1000 psi). The strains were

considered to be in equilibrium if they remained constant for six hours or more. The

experimental results showed that the three samples behaved in a somewhat similar fashion.

Figure 31 shows the linear and volumetric strains of sample #1, // and ┴ representing the

linear strains parallel and perpendicular to bedding planes, respectively. As expected, the

dimensions of the coal decreased with increasing helium pressure in all three directions but

they were not identical. The two measured linear strains in direction parallel to bedding

(//) exhibited almost the same deformation, as illustrated in Figure 31. The strain in

Page 112: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

97

direction perpendicular to bedding (┴) was slightly smaller than // at each pressure step.

This may be the result of anisotropy of the elastic property of coal. The stiffness of coal in

direction perpendicular to bedding is expected to be greater than in the direction parallel to

bedding since the coal cleats are vertically, or near vertically, oriented, resulting in the

elastic modulus in the direction parallel to bedding planes being smaller than in the

perpendicular direction. This is conceptually explained by Figure 1(b).

The volumetric strains of three samples were plotted versus helium pressure, as shown

in Figure 32. The volume of coal matrix decreased linearly with helium pressure due to

compression of the solid coal grain. Based on Figures 31 and 32, one definitive conclusion

can be made that coal is an elastic porous media with non-absorbing gas for the applied

pressure range, that is, at or below 6.9 MPa (1000 psi). Two evidences support this

conclusion: (1) both strains, linear and volumetric, returned to near-zero after helium

depletion, as shown in Figure 31, and (2) a linear stress/pressure-strain relationship is

observed in Figure 32. In other words, the mechanical induced strain by helium was

reversible. This finding also provided evidence that the mechanical compression induced

strain can be described by Hooke’s law in the strain models, which has been an assumption

in previous theoretical studies 18,20.

3.2 CO2 injection results

Following the helium cycle, Sample #1 was flooded with CO2 to estimate the CO2

sorption-induced deformation. Figure 33 shows the volumetric and three linear strains

induced by CO2 injection. A few features are apparent in Figure 33. First, the sorption

induced linear strains were anisotropic. The strain perpendicular to the bedding plane

Page 113: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

98

direction (┴) was greater than that parallel (//) to it. Generally, // was 55 to 70% of ┴ at

each pressure step. Second, the two parallel strains were almost same, qualitatively as well

as quantitatively. Third, the strain induced by adsorption/desorption of CO2 was not fully

reversible. Also, the amount of residual strain after depletion was not the same in the three

directions. The irreversible strain parallel to bedding was greater than that perpendicular to

it, as shown in Figure 33. Fourth, the time required for mechanical equilibrium was

significantly shorter than that for sorption equilibrium. For each injection step, the strain

initially decreased due to compression and then gradually increased due to the sorption

effect since ad-/de- sorption is a slow process compared to mechanical compression/

decompression. Lastly, it was observed that the sorption induced strain at each CO2

injection pressure shows a slow decrease after the peak strains. This strain decrease might

be attributed to the slow gas absorption in the coal matrix and the absorption effect leads

to a net decrease of equilibrium pressure over the testing time scale and which will decrease

the sorption induced strain as shown in Figure 5. This phenomenon was also confirmed by

the methane adsorption induced strain profiles as shown in Figure 6.

3.3 Methane injection results

Methane was injected into Samples #2 and #3 in six steps to a final pressure of 6.9

MPa (~1000 psi). The results suggested that the two samples followed the same pressure

and strain relationship. Methane was then bled out from Sample #2. The evolution of

strains, both volumetric and linear, is shown in Figure 34 for sample #2. Similar to CO2

injection, the methane induced strains were anisotropic and the // was about 40 to 60%

that of ┴. The experimental results, once again, showed that the two strains parallel to

Page 114: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

99

bedding were almost identical. After methane was fully depleted, the sorption induced

linear strains were almost reversible in each direction, as shown in the bottom right corner

in Figure 34. The irreversible portion is less than 3% of the total induced strain in direction

parallel to bedding and approximately 1% for direction perpendicular to bedding. Taking

into account the experimental error and strain gauge accuracy (± 0.5%), we can conclude

that sorption induced strain is reversible for methane.

3.4 Comparison of methane and CO2

The coal volumetric strains with exposure to methane and CO2 were plotted as a

function of gas pressure, as shown in Figure 35. For pressures up to 5.9 MPa (~850 psi),

the volume of coal sample increased by ~1.87% with CO2 injection. For methane injection,

the volume of coal sample increased by ~0.6% at the final pressure of 6.9 MPa (1000 psi).

The swelling of coal with CO2 is more than three times as that induced by methane, which

is consistent with the previous laboratory observations 33,40.

The injection (methane and CO2) induced volumetric strains shown in Figure 35

clearly exhibit a strong similarity with typical sorption isotherms. Hence, the measured

volumetric strains were fitted to a model similar to the Langmuir sorption model, initially

proposed by Levine 32:

휀𝑉 =𝜀𝑚𝑎𝑥𝑃

𝑃+𝑃𝜀 (23)

where, εV is the induced volumetric strain at pressure P, and εmax and Pε are the two

Langmuir-type constants. The equation provides an empirical measure to calculate the

sorption induced strain at any given pressure. It should be noted that this is only valid for

low pressure ranges, where the adsorption-induced strain dominates the volumetric

Page 115: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

100

behavior 20. In the current study, this relationship is valid since the maximum gas pressure

is ~6.9 MPa (1000 psi). The modeled results, along with modeled equations, are shown in

Figure 36. The figure shows that the modeled volumetric strains match well with the

measured values for both CO2 and methane. The Langmuir-type constants were estimated

to be εmax = 0.0257 and Pε = 2.15 MPa (312 psi) for CO2 and εmax = 0.0096 and Pε = 4.38

MPa (635 psi) for methane.

3.5 CO2 displacement results

The resulting variation in volumetric strain for sample #3 is shown in Figure 37. As

expected, the sample subject to CO2 injection exhibited a significant increase in strain,

almost three times that induced by methane at 5.9 MPa (850 psi). In this case, the induced

strain (1.68%) after methane was totally displaced by CO2 is slightly smaller than that

induced by pure CO2 injection at 5.9 MPa (850 psi), which was ~1.87%.

The sorption induced strain follows the Langmuir-type fit for pure gas (methane and

CO2), as discussed above. The critical question now was developing a means to describe

the gas mixture induced strain. The extended Langmuir model has been used to model and

predict the mixture gas adsorption on coal 42,43 and was believed to be appropriate at this

stage.

An analysis of whether the extended Langmuir-type model can successfully describe

the mixture gas induced strain or not was carried out. In order to model the binary gas

induced strain, pure gas induced strain data is required. The Langmuir-type model was

used to calculate induced strain for each gas component. In general, the total volumetric

strain induced by n-component gas mixture can be calculated as a sum of each gas-

component induced strain, mathematically given by:

Page 116: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

101

휀𝑉 = ∑ 휀𝑖𝑛𝑖=1 (24)

where, i is sorption-induced strain for gas component i and can be calculated as follows:

휀𝑖 =𝜀max (𝑖)𝑃𝑖𝑏𝑖

1+∑ 𝑃𝑖𝑏𝑖𝑛𝑖=1

(25)

where, max(i) and bi (equals to 1/Pε) are the Langmuir-type fit parameters for gas component

i and Pi is the corresponding gas partial pressure of component i.

In the present study, the gas mixture was two-component and the total volumetric strain

induced by methane and CO2 mixture is given as follows:

휀𝑉 =𝜀max (𝐶𝐻4)𝑃𝐶𝐻4𝑏𝐶𝐻4+𝜀max (𝐶𝑂2)𝑃𝐶𝑂2𝑏𝐶𝑂2

1+𝑃𝐶𝐻4𝑏𝐶𝐻4+𝑃𝐶𝑂2𝑏𝐶𝑂2 (26)

The Langmuir-type strain constants obtained by fitting the experimental results of

previous section were estimated and these are listed in Table 6. At each displacement step,

the partial pressure of methane was calculated from the methane molar fraction using GC

measured gas concentration. By substituting these parameters into equation (4), the

extended Langmuir-type modeled strains were calculated. There are presented in Figure

38. The model predicted perfectly the volumetric strain at low CO2 concentration, which

was 30% or less. It overestimated the gas mixture induced volumetric strain when CO2

concentration was greater than 30%, as shown in Figure 38. The discrepancy increased

with increase in CO2 concentration. The value predicted by extended Langmuir model is

12% more than the measured value at 100% CO2. This strain discrepancy between modeled

and measured results may be attributed to two reasons. First, the irreversible methane

sorption may occur due to methane-CO2 counter-diffusion effect. When the CO2 was

injected into methane-saturated coal, the methane molecules need to diffuse out from the

internal pores and CO2 molecules need to diffuse into the pores to achieve the

Page 117: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

102

multicomponent gas sorption equilibrium. This counter-diffusion phenomenon potentially

causes the irreversible methane adsorption in the pores because methane molecules are

more difficult to diffuse out in the methane-CO2-coal system than the pure methane gas

environment. Second, the micro-pore accessibility may be modified by gas mixture

injection 44. With continuous CO2 injection, a small portion of open pore might be closed

to CO2 at high CO2 concentration due to the coal softening effect. Further investigations

are required to test these two hypotheses.

4. Influence of anisotropic deformation on permeability during CO2 injection

Since deep coal is a potential candidate formation for carbon sequestration, the

permeability variation with CO2 injection is discussed further. A number of analytical

permeability models for coal permeability have been derived to incorporate the impact of

sorption-induced deformation 9,13,14,22,32. However, all these models assumed that the coals

are isotropic in both mechanical properties and sorption-induced strain. For coals with little

or small anisotropic sorption-induced strain behavior, the assumption of isotropy is a

reasonable approximation. However, this may not be the case for coal types used in

previous 15,19,32,33 and current laboratory studies. The assumption of isotropic sorption-

induced strain may, therefore, result in erroneous prediction of coal permeability.

Therefore, the quantitative influence of anisotropic deformation on coal permeability must

be considered and addressed although it is a challenge to incorporate anisotropic

deformation into account for most of the models 9,14,30,22,32 given that the derivation of these

models is based on overall sorption induced volumetric strain. In order to address this, a

permeability model with directional linear strain is required. Ma et al. 12 proposed a

screening model which included linear strain. The model was selected to investigate the

Page 118: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

103

impact of anisotropic deformation on permeability variation. The basis of the model was

constant reservoir volume and is rewritten as:

𝑘

𝑘0=

(1+2

𝜙0∙(𝜀𝑙−𝑠𝑜𝑟𝑝+

1−𝜈

𝐸∙(𝑝−𝑝0)))

3

1−𝜀𝑙−𝑠𝑜𝑟𝑝−1−𝜈

𝐸∙(𝑝−𝑝0)

(28)

where, k and k0 are the permeability at reservoir pressure p and p0, respectively, 0 is the

initial cleat porosity, E is Young’s modulus, is Poisson’s ratio, 휀𝑙−𝑠𝑜𝑟𝑝is pure sorption

induced linear horizontal strain for a single matchstick. Originally, 휀𝑙−𝑠𝑜𝑟𝑝 was calculated

by the following equation with isotropic assumption [12]:

휀𝑙−𝑠𝑜𝑟𝑝 = −1 + √1 + 휀𝑙 ∙ (𝑝0

𝑃𝜀+𝑝0−

𝑝

𝑃𝜀+𝑝) (29)

where, l and P are Langmuir type strain constants for the pure sorption induced

volumetric strain. These two constants are different from the direct laboratory measured

constants described in equation (1), since the measured strains shown in Figures 31 and

32 include two components, that is, pure sorption induced strain and mechanical

compression strain, respectively 20. The pure sorption induced strains were calculated by

subtracting the mechanical compression strain, a negative quantity, from the methane/CO2

injection measured strain. The pure sorption induced volumetric strain results for methane

and CO2 are shown in Figure 39. Using Langmuir type curve fitting, the l and P were

regressed to be 0.0278 and 2.21 MPa (320 psi). Thus, 휀𝑙−𝑠𝑜𝑟𝑝 was modeled as a function

of CO2 injection pressure.

Under constant volume boundary condition, there are no “overall” horizontal and

vertical strains because of the fixed boundary. However, for a single matchstick, shown in

Page 119: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

104

Figure 40, the localized horizontal strain is allowed, which tends to decrease the cleat

aperture resulting from the CO2 induced swelling strain. To satisfy the zero vertical strain,

a single matchstick deformation in vertical direction is not permitted and it will transfer to

horizontal strain though the Poisson effect, under conditions of elasticity, will hold. Here

we define it as “pushback” effect as shown in Figure 40. Conceptually, the single

matchstick initially swells in both vertical and horizontal directions, and then the vertical

deformation is pushed back to its pre-injection condition. Thus, the height of each

matchstick (h) maintains its original value. As illustrated in Figure 40, 휀𝑙−𝑠𝑜𝑟𝑝 is given by:

휀𝑙−𝑠𝑜𝑟𝑝 =𝑎−(𝑎+∆𝑎+𝑎′)

𝑎= −

∆𝑎

𝑎−

𝑎′

𝑎 (30)

where, −∆𝑎

𝑎 is purely sorption-induced horizontal strain and −

𝑎′

𝑎 is “pushback” induced

horizontal strain. The −∆𝑎

𝑎 value was directly calculated from experimental data by

subtracting the linear helium compression strain from the measured linear strain with CO2

injection. Since behavior in the two horizontal directions is almost the same, average of the

two was calculated and plotted in Figure 41. The pure sorption-induced linear strain was

then fitted using Langmuir model and the two strain constants were regressed to be 0.0073

and 1.9 MPa (275 psi). −𝑎′

𝑎 was obtained by multiplying Poisson’s ratio ( with the

vertical linear sorption-induced strain alone. Figure 41 also presents the pure vertical linear

sorption-induced strain along with the Langmuir fit. The two Langmuir constants were

0.0114 and 1.7 MPa (246 psi), respectively.

Substituting equations (6) and (7) into equation (5) yields the permeability variations

with CO2 injection for isotropic and anisotropic swelling, respectively. Since the primary

Page 120: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

105

objective for the permeability discussion is to investigate the impact of the anisotropic

directional sorption induced strain on permeability changes, other required parameters are

assumed to reasonable values based on the published literature. All parameters are listed

in Table 7. Based on the modeling exercise, the results are shown in Figure 42. It can be

seen that the permeability loss with CO2 injection was overestimated under isotropic

swelling assumption. This is expected since the vertical sorption linear swelling strain is

considered to be solely responsible for the permeability loss under isotropic assumption.

However, only partial vertical strain was transferred to horizontal strain under anisotropic

swelling as a result of the in situ “pushback” condition. Another important finding is that

permeability rebound occurs much earlier for in situ “pushback” condition than under

isotropic swelling assumption. This might possibly explain the CO2 injectivity increase in

the Yubari injection site in Japan where, when the injection pressure was kept constant at

15 MPa, the CO2 plume gradually migrated away from wellbore and led to an overall

effective permeability increase 45. The increase of CO2 injectivity in this site might also be

attributed to the CO2 injection induced micro-fracturing. When the coal was saturated with

CO2 at 15 MPa, the swelling strain will induce excess stresses due to the confined

geological boundary 40,46-48. This sorption-induced excess stresses might exceed the

compressive strength of coal to cause the micro-fracturing resulting in an overall increase

of fluid injectivity.

To summarize, this discussion indicates that the anisotropic effect of sorption induced

strain does influence the permeability change profile and proper consideration of

anisotropic swelling may increase the accuracy of permeability prediction during CO2

injection. Further study is required to apply the anisotropic sorption-induced strain into

Page 121: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

106

other boundary condition, such as, uniaxial strain which is the most accepted one when

modeling coal permeability.

5. Conclusions and Summary

This study has addressed the coal matrix deformation behavior with helium, methane,

CO2 and methane/CO2 flooding. Our aim was to measure the sorption-induced strains of

coal and to determine the anisotropy of deformation. The orthogonal linear strains for

Illinois coal were experimental measured for different gases. With anisotropic

shrinkage/swelling data, Ma et al. coal permeability model was employed to quantify the

impact of anisotropic deformation on permeability variations with CO2 injection. Overall,

our finding can be summarized as follows:

1. The mechanical induced compression linear strains in direction parallel to

bedding deform identically as each other, which are smaller than in the direction

perpendicular to bedding.

2. The sorption-induced linear strains are anisotropic for both methane and CO2.

Generally, the strain measured perpendicular to the bedding (┴) was greater than

parallel (//) to it.

3. The CO2 sorption-induced strains were not fully reversible, but they are for

methane.

4. The volumetric CO2 sorption induced strain was more than three times that

induced by methane.

5. With CO2 displacement, the induced strain (1.68%) after methane was completely

displaced by CO2, was slightly smaller than with CO2 injection, which was

Page 122: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

107

~1.87%. The extend-Langmuir type model can predict the strain behavior for CO2

concentration up to 30%. However, the model overestimated the strain for CO2

concentrations greater than 30%.

6. It was found that the application of isotropic deformation in permeability

prediction model can overestimate the permeability loss compared to anisotropic

deformation. This demonstrates that the anisotropic coal deformation should be

further considered to predict the permeability behavior in the coal-CO2

sequestration project.

References

Busch, A., Gensterblum, Y., 2011. CBM and CO2-ECBM related sorption processes in

coal: a review. International Journal of Coal Geology 87, 49-71.

Clarkson, C. R., Bustin, R. M., 2000. Binary gas adsorption/desorption isotherms: effect

of moisture and coal composition upon carbon dioxide selectivity over methane.

International Journal of Coal Geology 42, 241-271.

Connell, L.D., Detournay, C., 2009. Coupled flow and geomechanical processes during

enhanced coal seam methane recovery through CO2 sequestration. International

Journal of Coal Geology 77, 222-233.

Cui, X., Bustin, R. M., 2005. Volumetric strain associated with methane desorption and

its impact on coalbed gas production from deep coal seams. AAPG 89, 1181-1202.

Day, S., Fry, R., Sakurovs, R., 2008. Swelling of Australian coals in supercritical CO2.

International Journal of Coal Geology 74, 41-52.

Day, S., Fry, R., Sakurovs, R., 2012. Swelling of coal in carbon dioxide, methane and

their mixture. International Journal of Coal Geology 93, 40-48.

Dutta, P., Harpalani, S., Prusty, B. K., 2008. Modelling of CO2 sorption on coal. Fuel 87,

2023-2036.

Page 123: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

108

Gunter, W. D., Mavor, M. J., Robinson, J. R., 2005. CO2 storage and enhanced methane

production: field testing at Fenn Big Valley, Alberta, Canada, with application.

Proceedings of the 7th International conference on Greenhouse Gas Control

Technologies: Volume I, Rubin, E. S., Keith, D. W., Gilboy, C. F. (Eds.) Elsevier

Ltd., 413-421.

Gunter, W. D., Wong, S., Law, D., Feng, S., Ye, J., Fan, Z., 2005. Enhanced coalbed

methane (ECBM) field test at south Qinshui Basin, Shanxi Province, China. GCEP

Workshop, August 22-23, Beijing China.

Harpalani, S., Chen, G., 1997. Influence of gas production-induced volumetric strain on

permeability of coal. International Journal of Geotechnical and Geological

Engineering 15, 303-325.

Harpalani, S., Mitra, A., 2010. Impact of CO2 injection on flow behavior of coalbed

methane reservoirs. Transport in Porous Media 82, 141-156.

Harpalani, S., Prusty, B. K., Dutta, P., 2006. Methane/CO2 sorption modeling for coalbed

methane production and CO2 sequestration. Energy & Fuels 20(4), 1591-1599.

Kelemen, S. R., Kwiatek, L. M., 2009. Physical properties of selected block Argonne

Premium bituminous coal related to CO2, CH4, and N2 adsorption. International

Journal of Coal Geology 77, 2-9.

Koperna, G. J., Oudinot, A. Y., McColpin, G. R., Liu, N., Heath, J. E., Wells, A., Young,

G. B., 2009. CO2-ECBM/storage activities at the San Juan basin’s Pump Canyon test

site. SPE Annual Technical Conference and Exihbition October 4-7, New Orleans,

Louisiana, USA.

Langmuir, I., 1918. The adsorption of gases on plane surfaces of glass, mica, and

platinum. J. Am. Chem. Soc., 40, 1361-1403.

Laubach, S. E., Marrett, R. A., Olson, J. E., Scott, A. R., 1998. Characteristics and origins

of coal cleat: a review. International Journal of Coal Geology 35, 175-207.

Levine, J. R., 1996. Model study of the influence of matrix shrinkage on absolute

permeability of coal bed reservoirs. In: Gayer, R., Harris, I. (Eds.), Coalbed Methane

and Coal Goelogy, Geological Society Special Publication No 109 London, 197-212.

Liu, H. H., Rutqvist, J., 2010. A new coal-permeability model, internal swelling stress

and fracture-matrix interaction. Transport in Porous Media 82, 157-171.

Liu, J., Chen, Z., Elsworth, D., Miao, X., Mao, X., 2011. Evolution of coal permeability

from stress-controlled to displacement controlled swelling conditions. Fuel 90, 2987-

2997.

Liu, S., Harpalani, S., 2012. Gas production induced stress and permeability variations in

coalbed methane reservoirs. The 46th US Rock Mechanics/Geomechanics

Symposium Chicago, IL, June 24-27.

Liu, S., Harpalani, S., 2013a. A new theoretical approach to model sorption-induced coal

shrinkage or swelling. AAPG bulletin v. 97(7), 1033-1049.

Page 124: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

109

Liu, S., Harpalani, 2013b. Permeability prediction of coalbed methane reservoir during

primary depletion. International Journal of Coal Geology 113, 1-10.

Liu, S., Harpalani, S., 2014a. Evaluation of in situ stress changes with gas depletion of

coalbed methane reservoirs. Journal of Geophysical Research – Solid Earth, v.

119(8), pp 6263-6276.

Liu, S., Harpalani, S., 2014b. Compressibility of sorptive porous media: part II –

experimental study on coal, AAPG Bulletin, v. 98, No. 9, pp. 1773-1788.

Liu, S., Harpalani, S., 2014c. Determination of the effective stress law for deformation in

coalbed methane reservoirs, Rock Mechanics and Rock Engineering, v. 47(5), pp.

1809-1820.

Ma, Q., Harpalani, S., Liu, S., 2011. A simplified permeability model for coalbed

methane reservoir based on matchstick strain and constant volume theory.

International Journal of Coal Geology 85, 43-48.

Majewska, Z., Majewski, S., Zietek, J., 2010. Swelling of coal induced by cyclic

sorption/desorption of gas: experimental observations indicating changes in coal

structure due to sorption of CO2 and Ch4. International Journal of Coal Geology

83(4), 475-483.

Moffat, D. H., Weale, K.E., 1955. Sorption by coal of methane at high-pressures. Fuel

34, 449-462.

Mora, C. A., Wattenbarger, R. A., 2009. Analysis and verification of dual porosity and

CBM shape factors. Journal of Canadian Petroleum Technology 48, 17-21.

Ohga, K., Fujioka, M., Yamaguchi, S., Nako, M., 2006. Field test of CO2 sequestration

and ECBM in Japan. International Coalbed Methane Symposium, Tuscaloosa,

Alabama, Paper 0635.

Ong, S., Zheng, Z., Chhajlani, R., 2001. Pressure-dependent pore volume compressibility

– a cost effective log-based approach. SPE Asia Pacific Improved Oil Recovery

Conference, Kuala Lumpur, Malaysia, 8-9 October, paper SPE72116.

Oudinot, A. Y., Koperna, G. J., Philip, Z. G., Liu, N., Heath, J. E., Wells, A., Young, G.

B., Wilson, T., 2011. CO2 injection performance in the Fruitland coal Fairway, San

Juan basin: results of a field pilot. SPE Journal 16(4), 864-879.

Palmer, I., Mansoori, J., 1998. How permeability depends on stress and pore pressure in

coalbeds: a new model. SPE Reservoir Evaluation & Engineering 1(6), 539-544.

Palmer, I., Mavor, M., Gunter, B., 2007. Permeability changes in coal seams during

production and injection. International Coalbed Methane Symposium, Tuscaloosa,

Alabama, Paper 0713.

Pan, Z., Connell, L. D., 2007. A theoretical model for gas adsorption-induced coal

swelling. International Journal of Coal Geology 69, 243-252.

Page 125: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

110

Pan, Z., Connell, L. D., 2011. Modelling of anisotropic coal swelling and its impact on

permeability behavior for primary and enhanced coalbed methane recovery.

International Journal of Coal Geology 85, 257-267.

Reeves, S., Oudinot, A., 2005. The Allison Unit CO2-ECBM pilot – a reservoir and

economic analysis. International Coalbed Methane Symposium. Tuscaloosa,

Alabama, Paper 0522.

Robertson, E. P., 2008. Christiansen, R. L., A permeability model for coal and other

fractured, sorptive-elastic media. SPE Journal 13, 314-324.

Seidle, J. P., Jeansonne, M. W., Erickson, D. J., 1992. Application of matchstick

geometry to stress dependent permeability in coals. SPE Rock Mountain Regional

Meeting, Casper, Wyoming, May 18-21.

Shi, J. Q., Durucan, S., 2005. A model for changes in coalbed permeability during

primary and enhanced methane recovery. SPE Reservoir Engineering 8, 291-299.

Silva, P. N. K. D., Ranjith, P. G., Choi, S. K., 2012. A study of methodologies for CO2

storage capacity estimation of coal. Fuel 91, 1-15.

US EIA, 2013. Natural gas gross withdrawals and production, Released on 4/30/2013. US

Energy Information Administration.

Wang, S., Elsworth, D., Liu, J., 2012. A mechanistic model for permeability evolution in

fractured sorbing media. Journal of Geophysical Research 117, B06205.

Wageningen, van, W. F. C., Maas, J. G., 2007. Reservoir simulation and interpretation of

the RECOPOL ECBM pilot in Poland. International Coalbed Methane Symposium,

Tuscaloosa, Alabama, Paper 0702.

Warren, J. E., Root, P. J., 1963. The behavior of naturally fractured reservoirs. SPE

Journal 3, 245-255.

Yang, R. T., 1987. Gas separation by adsorption processes. Butterworth Publisher,

Stoneham MA, USA.

Zarebska, K., Ceglarska-Stefanska, G., 2008. The change in effective stress associated

with swelling during carbon dioxide sequestration on natural gas recovery.

International Journal of Coal Geology 74, 167-174.

Zhang, R., Liu, S., Bahadur, J., Elsworth, D., Melnichenko, Y., He, L., Wang, Y., 2015.

Estimation and modeling of coal pore accessibility using small angle neutron

scattering. Fuel 161, 323-332.

Page 126: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

111

Appendix

Figure 29 Conceptual coal cleat system: (a) cubic geometry; (b) matchstick geometry (after Pan

and Connell, 2012)

(b) (a)

Page 127: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

112

Figure 30 Schematic of the experimental setup to measure the sorption induced strain with gas

mixture.

Figure 31 Helium injection and depletion induced strains.

Page 128: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

113

Figure 32 Volumetric strain with helium injection for three coal samples.

Page 129: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

114

Figure 33 CO2 injection and depletion induced strains.

Page 130: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

115

Figure 34 Methane injection and depletion induced strains.

Figure 35 Volumetric strain comparison of CO2 and methane injection and depletion

Page 131: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

116

Figure 36 Experimental strain results along with Langmuir-type model.

Figure 37 Volumetric strain by methane displacement by CO2.

Page 132: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

117

Figure 38 Volumetric strain by methane displacement by CO2 with extended Langmuir modeled

results.

Figure 39 Pure sorption induced volumetric strain by CO2 along with Langmuir type model.

Page 133: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

118

Figure 40 Conceptual deformation for a single matchstick under constant volume assumption.

Figure 41 Pure sorption induced linear strains by CO2 along with Langmuir type model.

Page 134: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

119

Figure 42 Comparison of permeability behavior for isotropic and anisotropic deformation

conditions.

Table 6 Volumetric strain Langmuir-type fit parameters for methane and CO2.

max(CH4)

Dimensionless

bCH4(1/P(CH4)),

MPa-1

max(CO2)

Dimensionless bCO2(1/P(CO2)),

MPa-1 0.0096 0.228 0.0257 0.465

Table 7 Input parameters for Ma et al. permeability model.

0 (%) E (MPa) Isotropy

Anisotropy

Horizontal Vertical

l P(MPa) l P(MPa) l P(MPa)

2* 2000* 0.25* 0.0278 2.21 0.0073 1.19 0.0114 1.7

* Value assumed.

Page 135: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

120

Chapter 5 Apparent Permeability Characterization and Multiple Flow

Regimes Modeling with Helium and CO2 Injection under Hydrostatic

Condition

Abstract

Gas transport mechanism in shale is critical for the prediction of shale permeability

and gas production. Gas transport in shale matrix no longer always falls into the continuum

flow regime described by Darcy’s law because of existing of multiscale pores from nano-

to macro-scale. A multi-mechanistic flow process is always happening during shale gas

production, including Darcy viscous flow, slip flow, transition flow and Knudsen diffusion

and their proportional contributions to apparent permeability are constantly changing with

pressure evolution. In this study, several cycles of laboratory measurements of shale

permeability with helium and CO2 injections were carried out under hydrostatic condition

with different confining stresses by using pulse-decay method. CO2 apparent permeability

can be higher than He at pressure higher than 1000 psi, which may result from limited shale

adsorption capacity. Helium permeability is more sensitive to the variation of Terzaghi

stress than CO2 and it is independent of pore pressure. The true effective stress coefficient

can be found two values at low pressure region (<500 psi) and high pressure region (>500

psi). The negative value indicates Knudsen diffusion and slip flow effect have more impact

on apparent permeability than Terzaghi stress. Knudsen diffusion was proven to be

significant at low pressure/late-time shale gas production in both laboratory and modeling

aspects. Knudsen diffusive permeability can be significantly varied along with average

pore size of shale. However, dynamic flow regimes, effective stress and sorption-induced

matrix/pore deformation are critical in shale reservoir characterization and shale gas

Page 136: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

121

flow/permeability modeling.

1. Introduction

In recent years, gas shale reservoirs play important roles in natural gas supply in the

United States. The most recent annual energy report by US Energy Information

Administration (US-EIA) reveals an incredible increase of shale gas production from 2007

to 2014, representing up to 40% of total natural gas production (EIA, 2015). Generally, the

hydraulic fracturing stimulation is very effective to create complex fracture networks that

can connect the shale formation to the horizontal boreholes. However, the low matrix

permeability retards the gas migration from the tight matrix into fracture networks.

Dominant pore diameters of shale matrix can be smaller than 2 nm, which is significantly

smaller than average pore size in conventional sandstone and carbonate reservoirs (S. Liu,

2012; Swami, Clarkson, & Settari, 2012). Shale matrix associated with organic matters and

clay minerals has high gas storage capacity but limited gas deliverability because it is

extremely tight with very low diffusivity and/or permeability. The nature of tight structures

of shales introduces various flow dynamics which make the assessment of gas transport

properties is challenging and complex. It was widely accepted that gas transport in shale

reservoirs is a multi-mechanistic process, including both Darcy and non-Darcy flows

components including Knudsen diffusion, transition flow and slip flow (Javadpour, 2009;

Y. Li et al., 2014; Wu, Li, Wang, Chen, & Yu, 2015). And the dominant shrinkage/swelling

response of organic components of shale, due to the interaction with sorbing gases, causes

significant permeability variation (H. Kumar, Elsworth, Mathews, & Marone, 2015). Those

characteristics generally make laboratory assessment and modeling of shale permeability

Page 137: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

122

more complicated. Specifically, the extremely tight structure for intact shale core sample

without apparent fractures has no obvious gas flows under high stress conditions. As

expected, if a long intact shale core was used for the permeability measurement, it will take

incredibly long lab cycles for equilibrium. In addition, the measured permeability values

by using the classic analytical solution for permeability calculation may be not

representative since only Darcy’s law is assumed to be valid. Nano-pore and micro-pore

gas flow along with sorption effect need to be separately characterized and integrated to

account for the multi-mechanistic shale gas flow process.

In this study, two well-cut and polished shale thin disk samples will be used to conduct

the permeability measurements by using transient method (pulse-decay technique) in order

to characterize the shale gas permeability behavior under different conditions. Different

flow regimes will be separately analyzed in aspect of their contribution to total permeability

evolution. Other important effects, such as effective stress and sorption-induced

deformation, which were reported frequently in unconventional rock studies, will also be

considered as combined effects. They will be investigated together with the changing of

flow regimes in shale gas permeability measurements and real production processes.

Especially, to investigate the producing potential of late-time shale gas production period,

we will also conduct modeling study under the circumstance only Knudsen diffusion

dominates.

2. Background and Theories

1.1 Shale gas flow regimes

During shale gas reservoir production, Darcy flow permeability is still a prevailing

Page 138: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

123

term to quantify the deliverability of gas flow in shale fractures and matrix. However,

diffusion flow is just as important as Darcy flow in unconventional gas production

prediction. Figure 43 shows different flow mechanisms during gas shale reservoir

depletion. The average monthly production for 15 wells in Eagleford Shale formation

shows that the peak production happens in the late second month after the fracturing

treatment. Followed by a rapid decline till 20th month and then have a production plateau

at relatively low production rate. During early-time production, fracture (Darcy) flow

dominates since free gas molecules flowing in hydraulic fractures (large diameter flow

channels) occupy the flow path and driven by pressure gradient. The production reaches

the peak very quickly due to the fracture flow. Followed with the peak production, the rapid

decline happens due the quick depletion of the fractures. Although, desorption could

involve during the pressure decline stage, only a negligible portion of gas contribute to the

production because the slow diffusion process. Generally, diffusion flow at this phase

doesn’t really dominate total gas production except the free gas in fractures evacuate

quickly. Gas transport mechanism moves from continuum flow to a transition zone between

Darcy flow and Knudsen diffusion. This transition zone includes slip flow (0.001<𝐾𝑛<0.1)

and transition flow (0.1<𝐾𝑛<1) where neither continuum flow nor free molecular flow are

valid (Heller et al., 2014). gas flow in this zone can be considered a combination of Darcy

flow and Knudsen diffusion. After the fractures were depleted, the late-time production

period comes, when the reservoir fracture pressure becomes very low. At this stage, even

though the fracture permeability is high enough, the gas mass influx from matrix to

fractures becomes the controlling flow mechanism. This mass influx from matrix to

fracture is a multi-mechanical micro/nano-scale transport. Methane molecules firstly

Page 139: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

124

desorb from the internal surfaces of matrix into nano-pore system due to pressure driving

force and become free gas (Do, 1998). During desorption process, the distance between the

gas shale surface molecule layer and the adjacent molecules decreases, the effective surface

energy on shale kerogen surface also increases (S. Liu & Harpalani, 2014a). As a result,

kerogen shrinkage will also occur in shale matrix and enhance the nano-pore flow path for

gas molecules (Figure 44). Most of gas molecules start to transport through the nano-pore

by means of diffusion flow. Because of the wide range of pore size inside the shale matrix,

it can be expected a domination of diffusive flow or a combination of multiple flow regimes.

This diffusion dominate flow can sustain for decades which contributes for the late

production flat tail. Thus the gas flow permeability evolution at late production stage is

critical and significantly influence the late production behaviors (Y. Wang & Liu, 2016). In

conclusion, shale gas transport during production always includes all flow regimes (Darcy,

slip flow, transition flow and diffusion with gas desorption) in every stage, but their

proportional contribution will change with decreasing pressure. Due to the extremely small

pore size and/or volume in shale formations and complicated matrix shrinkage/swelling

effect, conventional Darcy flow and diffusion model cannot solely describe gas flow

transport in shale formations. It commonly requires to involve multi-mechanic flows

mainly including Darcy viscous flow and Knudsen diffusion (Javadpour, 2009). In other

words, a multi-mechanistical method that contains both breakdown analysis and optimized

integration is needed.

The breakdown of different flow regimes, such as Darcy flow, slip flow, transition

flow can be classified and distinguished by Knudsen number (𝐾𝑛) which is known as a

Page 140: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

125

function of pore size and pore pressure:

𝐾𝑛 =𝜆

2𝑟 (31)

where, r is pore radius and 𝜆 is gas mean free path defined as:

𝜆 =𝑘𝐵𝑇

√2𝜋𝑑𝑚2𝑝

(32)

where p is pore pressure and 𝑘𝐵 is the Boltzmann constant (1.3805 x 10-23 J/K).

The relationship between 𝐾𝑛, pore size and pressure is also plotted in Figure 45. The

pressure range follows the experimental pressure range that we used for the subsequent

laboratory study. For Marcellus shale reservoirs, since pore radius is in nanometers and

micrometers (C. R. Clarkson, Solano, & Bustin, 2013), 𝐾𝑛 is generally very low (𝐾𝑛 < 0.01)

(Karniadakis et al., 2005). The mean free path is negligible compared to extremely small

pore throat size. The flow is thus governed by viscous flow and driven by pressure gradient.

For this range of 𝐾𝑛, the intermolecular collisions are important and molecules’ collisions

with pore walls are negligible (Swami et al., 2012). Transition flow and slip flow

potentially can be modeled as a weighted combination of pure Poiseuille flow (Darcy) and

Knudsen flow (Heller et al., 2014; Kuila et al., 2013). Within the pore size range given in

Figure 45, gas flow in shale matrix lies mostly within transition flow and slip flow region.

It can potentially involve Darcy flow regime if the pressure is much higher, or include

Knudsen diffusion when pressure is low. In order to get multi-mechanistic flows, pulse-

decay transient method was used to measure shale matrix apparent permeability in this

study. All measurements were conducted at relative low pressures (<1500 psi) to guarantee

2~3 flow regimes can be covered and then combined flow regimes should be applied to

describe the gas flow and permeability evolution.

Page 141: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

126

The rate of gas flow between the shale matrix and fracture surface is very slow

(Javadpour, 2009; Swami and Settari, 2012) and was believed less important from the point

of view of well production. Diffusion, however, plays an important role in multi-

mechanistic gas flow and diffusivity is a critical parameter of tight reservoir rocks that can

limit the production when the fracture permeability is order of magnitude higher than the

matrix. The diffusion process in shale reservoirs is described by Fick’s Law and is driven

by the concentration gradient at which 𝐾𝑛 is relatively high (𝐾𝑛> 10) (Karniadakis et al.,

2005). Fick’s Law is commonly used to depict the relationship between the diffusional

mass influx and concentration gradient by assuming that the mass flowrate across a surface

is proportional to the concentration gradient, area of the surface, and diffusion coefficient

for a solid/gas medium. The diffusion in the shale matrix depends upon the matrix structure

and pressure (Javadpour et al., 2007; Javadpour, 2009; Swami and Settari, 2012). In

Javadpour (2009), a capillary tube gas flow model was proposed for shale to account for

Knudsen diffusion and slip flow using the slip boundary condition, which was combined

into Darcy flow to describe the apparent permeability. Darabi et al. (2012) proposed a new

apparent permeability model considering Knudsen diffusion, surface roughness and slip

flow. Although these apparent permeability models combined Darcy/continuum flow, slip

flow, transition flow and Knudsen diffusion components, their flow contribution weighting

factors were always assumed to be constants. It was noted that these models were not

capable of describing multi-mechanistic gas transport phenomenon. However, they provide

the fundamental understanding to address Knudsen diffusion flow in predicting the

complex gas transport process in shale matrix. Given that pores in shale matrix are

multiscale, from nano- and microscales, the diffusive flow is critical and controls the

Page 142: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

127

production at low reservoir pressure.

Navier Stokes equation that Darcy’s law was theoretically derived from (Hubbert, 1957)

only works for the continuum flow regime which includes Darcy viscous flow and slip

flow region. It breaks down when the flow regime moves into the transition region and the

free molecular flow region (Swami et al., 2012). In this case, the characterization in

molecular flow level is necessary. The mass-flow rate of the free molecular regime was

modeled by Knudsen (1909). Roy et al. (2003) showed that Knudsen diffusion in nano-

pores can be written in the form of pressure gradient. Gas mass flux by diffusion with

negligible viscous effects in a nano-pore is described as:

𝐽𝐷 = 𝐷𝐾𝑀Δ𝑝

𝑅𝑇𝐿 (33)

where, 𝐽𝐷 is the Knudsen mass flux, Δ𝑝 is the pressure drop across the rock sample, L is

the length of the medium and M is molar mass of tested gas. Also, Knudsen diffusion

constant 𝐷𝐾 is defined as:

𝐷𝐾 =2𝑟

3√

8𝑅𝑇

𝜋𝑀 (34)

In this Knudsen diffusion term, the porosity/tortuosity factor (𝜙/𝜏) is also introduced

to model Knudsen flow through porous media (Javadpour et al., 2007; Swami & Settari,

2012). This factor is to conceptualize the porous medium as a medium that consists a

certain percentage of open pores (porosity) of the mean pore diameter and have a degree

of interconnection resulting in paths longer than a direct path (tortuosity) (Javadpour et al.,

2007). In addition, the fractal dimension of the pore surface is included to consider the

effect of pore-surface roughness on the Knudsen diffusion coefficient (M. Coppens, 1999;

Page 143: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

128

M. O. Coppens & Dammers, 2006). Surface roughness describes the local heterogeneity of

the pores. Increasing surface roughness leads to an increase in residence time of molecules

in porous media and a decrease in Knudsen diffusivity. And this fractal dimension of the

pore surface is a quantitative measure of surface roughness that varies between 2 and 3,

representing a smooth surface and a space-filling surface, respectively (M. O. Coppens &

Dammers, 2006). Then the modified Knudsen diffusion constant becomes:

𝐷𝐾′ = (𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏√

8𝑅𝑇

𝜋𝑀 (35)

where, 𝐷𝑓 is the fractal dimension of the pore surface, and 𝛿′ is the ratio of normalized

molecular size 𝑑𝑚 to local average pore diameter 𝑑𝑝 , yielding 𝛿′ = 𝑑𝑚/𝑑𝑝 . Then we

substitute Eqs. (5) into (3):

𝐽𝐷 = 𝐷𝐾′ MΔ𝑝

𝑅𝑇𝐿= (𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏

MΔ𝑝

𝑅𝑇𝐿√

8𝑅𝑇

𝜋𝑀 (36)

At low gas pressure with small pore size, collisions are mainly between molecules and

walls. Knudsen diffusion flux depends on pressure gradient can be obtained by Eqs. (6)

(Choi, Do, & Do, 2001; M. O. Coppens & Dammers, 2006; Roy et al., 2003). In order to

model the gas flow permeability from gas flow model, mass flux term needs to be converted

into permeability term. Note that the relationship between mass flux and permeability in

continuous flow model can be determined as (Javadpour, 2009; Sakhaee-Pour & Bryant,

2012; Wu, Li, Wang, Yu, & Chen, 2014):

𝐽 = 𝑘𝜌

𝜇∇𝑝 (37)

where k is gas permeability. By comparing Eqs. (6) and (7), the apparent permeability of

Page 144: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

129

Knudsen diffusion can be mathematically written as:

𝐾𝐷 = 𝐽𝐷𝜇

𝜌∇𝑝= (𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏

𝜇M

𝜌𝑅𝑇√

8𝑅𝑇

𝜋𝑀 (38)

1.2 Permeability measurement approaches

Pulse-decay technique is believed an effective method that can be used to measure the

permeability of tight and ultra-tight reservoir rocks (Brace et al., 1968; Hsieh et al., 1981;

Dicker and Smits, 1988; Jones, 1997; Cui et al., 2009; Wang et al., 2011; Wang et al., 2015;

Kumar et al., 2015). The advantages of this method is that the permeability can be

calculated directly from the linear portion of the solution.

In order to test the effect of effective stress on permeability, excluding the contribution

of flow regime effect, different confining stresses and pore pressures were applied in the

pulse-decay experiments. Terzaghi effective stress (or simple effective stress), as a function

of confining stress and pore pressure, can reflect the change of many quantifiable physical

properties in porous rocks (Terzaghi, 1943):

𝜎𝑇𝑒𝑟 = 𝜎𝑐 − 𝑝 (39)

where 𝜎𝑐 is the confining stress, i.e. hydrostatic stress in this case. Both confining stress

and pore pressure can be controlled and monitored so that it is able to conduct series of

relationships between Terzaghi stress, confining stress, pore pressure and permeabilities.

However, for permeability experiments, the effective stress law with an appropriate

coefficient has been used instead of Terzaghi’s principle (Heller et al., 2014; Kwon,

Kronenberg, Gangi, & Johnson, 2001; Zoback & Byerlee, 1975):

𝜎𝑒 = 𝜎𝑐 − 𝜒𝑝 (40)

where 𝜎𝑒 is the effective stress and 𝜒 is the effective stress coefficient for permeability. To

Page 145: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

130

calculate 𝜒, we use the ratio of slope method presented by Kwon et al. (2001):

𝜒 = −𝜕𝑙𝑜𝑔 𝑘/𝜕𝑝

𝜕𝑙𝑜𝑔 𝑘/𝜕𝜎𝑐 (41)

As a multiplier of pore pressure, 𝜒 determines the relative sensitivity of permeability

to the changes in 𝜎𝑐 and p. For example, when 𝜒 is smaller than 1, it indicates the variation

of pore pressure has less impact on permeability evolution than confining stress does.

Zoback and Byerlee (1975) proposed a conceptual matrix model for tight rocks, suggesting

that the compressibility of fluid flow pathway and granular rock framework can be different

in values which differentiates the impacts of 𝜎𝑐 and p. And Kwon et al. (2001) has found

that the variation of 𝜒 can result from the clay content. By fitting all the permeability data

to the effective stress law, we can address the proportional effects of 𝜎𝑐and p, respectively.

3. Experimental work

3.1 Sample Procurement and Preparation

With tight structure, the equilibrium time for pressure pulse-decay is extremely long.

This long equilibrium time hurts the efficiency of lab testing, and even no pressure drop

was observed under very high stress conditions after a few hours of pulse injection.

Therefore, the shale thin discs were preferred for the permeability measurements which

can significantly shorten the equilibrium time and this has been commonly used for shale

permeability measurements (Heller et al., 2014). In this study, Marcellus shale drilled cores

were thus prepared as thin disks for the gas apparent permeability tests. The prepared thin

disks were in a thickness range from 3 to 6 mm, 1 inch in diameter. All trimmed surfaces

were polished to enable proper placement in the triaxial cell. Two well-prepared shale

samples in different thickness were shown in Figure 46. After the shale sample disks were

Page 146: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

131

dried in oven, they were then preserved in a dry and clean plastic sample bag in a lab-use

alloy box for 3 hours before being put into triaxial cell.

3.2 Experimental stress boundary conditions

Hydrostatic condition was applied in this series of shale permeability measurements.

In triaxial cell test, the constant hydrostatic stress condition refers to the axial stress is equal

to the 𝜎𝑐 at a constant value throughout the course of each experiment duration. The

hydraulic stresses are generated and maintained at a constant value by computer-controlled

syringe pumps.

3.3 Experimental setup and procedure

Figure 47 shows the schematic of the pulse-decay experimental setup for the pulse-

decay permeability measurement. The setup includes a Temco triaxial cell core holder, two

ISCO syringe pumps, flowrate and pressure monitoring and recording systems. The syringe

pumps are capable of maintaining or changing stresses in a controlled manner to the desired

levels. Software panel, desktop LabVIEWTM (laboratory virtual instrument engineering

workbench) software control panel, is developed to accurately control the pumps and

record the stresses and injection/ejection volumes of fluid. A rubber jacket is used to isolate

the sample from the confining fluid. The experimental temperature was kept constant at

296 K. For each experiment, the shale thin disk was sandwiched between two porous disks.

Two loading platens on two sides of triaxial cell serve as seals and the source to apply the

axial stress. Two comparable fixed volumes, large diameter Swagelok tubing, serving as

the upstream and downstream gas reservoirs, are estimated to be 31989 mm3 and 19628

mm3. Two high-accuracy USB-based Omega pressure transducers were installed to

continuously monitoring and recording pressure-time responses with high sampling rate

Page 147: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

132

during the experimental measurements. The pressure sampling rate was set to 5 readings

per second which can give accurate and continuous pressure pulse-decay profile.

The apparent permeability measurements on two Marcellus shale samples in different

thickness were carried under two different hydrostatic stresses, with two different gases.

Each experimental condition was identically repeated once in case of hysteresis effect.

Helium (He) and CO2 were used as the testing fluids injected by another high-capacity

syringe pump. Both helium and CO2 were chosen because we want to make comparison

between soring and non-sorbing gases. Two hydrostatic stresses, 1600 and 3000 psi, were

applied alternatively for testing the influence of external stress and effective stress on shale

permeability evolutions. An entire view of experimental sequence is listed in Table 8. As

a non-sorbing gas, helium was injected to test the shale matrix flow property to exclude

gas sorption effect. The sample was gradually stressed to 1600 psi as the hydrostatic stress

by two syringe pumps. After the mechanical equilibrium was achieved, the entire system

was vacuumed by a vacuum pump to remove the residual air in the gas flow system. For

the very first step, helium was injected at 100 psi for upstream reservoir and downstream

remains vacuum. Then the valves between sample and downstream were opened to

discharge the gas through the sample to downstream reservoir. This small pulse of gas was

driven by the pressure difference between upstream and downstream. The equilibrium

pressure was self-finalized at 64 psi in this step. Sample permeability at this step was

estimated based on the pressure pulse-decay responses. To proceed with the second step,

both upstream and downstream pressure were increased to 500 and 300 psi, respectively.

This gave us an equilibrium pressure roughly around 400 psi. After several steps, gas

pressure almost reached hydrostatic stress which should not be exceeded by gas pressure.

Page 148: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

133

In order to ensure the repeatability of the experimental results, we repeated all the above

mentioned experimental steps to get another set of estimated permeability data. Then we

depleted and vacuumed system, and gradually increased the external stress from 1600 to

3000 psi. Same strategy except larger pressure spans, 70~1970 psi were applied in this

group of measurement. Next, in order to make comparison between gases, the external

stresses for CO2 were set as the same as the helium depletion, which are 1600 psi and 3000

psi, respectively. Similar injection steps were applied from 55~1200 psi and 60~1484 psi.

The other sample with 3.91 mm in thickness was measured with exactly same procedures.

All the permeabilities were calculated from the pressure response data recorded during

pulse-decay cycles, using Eqs. (12) and (13).

4. Results and discussion

The measured permeability results of helium and CO2 for two disk samples (two cycles

each) was shown in Figures 48, 49 and 50. According to Figures 48 and 49, every repeated

pair of experimental cycles has only minimum difference, or even same in permeability

values. Also, the permeability data between the two samples are extremely close to each

other, and their trends are extremely uniform. Although strong hysteresis has been reported

in recent shale core permeability measurements possibly due to fracture closure (Rydzy,

Patino, Elmetni, & Appel, 2016), our experimental data are very close in both values and

trends, and our measurements showed their good consistency and repeatability. This

repeatability is critical because it provide the precision for the basis of valid experimental

measurements. In shale permeability measurements, repeated results in permeability data

indicates the structure of shale sample matrix is not damaged by applying high 𝜎𝑐 and high

p injection/depletion. The pore structure and micro-fractures maintain their integrity. The

Page 149: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

134

consistency of data between samples different in length indicates the matrix structures for

the sample collected from the same source can be very similar. Consistent and accurate

results help us to justify the utility of collected experimental data and consolidate any

hypotheses and conclusions based on them.

4.1 Permeability evolution with flow regimes

For all groups of data, the permeabilities initially decline with pressure injection from

50 to 500 psi and then either start to make a slight or even sharp turn and increase quickly

with the increase of gas pressure. At low pressure, nano-/micro-scale flow, such as Knudsen

diffusion, take large place and has reduced contribution with increasing pressure. With

decreasing Terzaghi stress (in black dash line), shale gas permeability still has a decreasing

tendency. This is where late-time production happens in actual shale gas production and

apparently diffusion flow has significant influence on gas deliverability. Similar behavior

can also be observed in CO2 permeability curves. The reasonable explanation for this

phenomenon is that the tight shale matrix, full of nano-scale and micro-scale pores, and

low gas pressure limited the control of Darcy flow effect and made Naiver-stokes equation

invalid. The apparent permeability of gas increases when pressure drops, since Knudsen

diffusion can be more significantly at low pressure. With the non-sorbing gas helium, we

can claim this statement with more confidence because we totally exclude the sorption

effect on permeability variation in helium injection. At high pressure, macro-scale flow

like Darcy viscous flow, will become dominant and increase with the decrease of Terzaghi

stress as shown in both Figure 48 and 49. If the gas pressure reaches a higher level, we

may see gas permeability values reach an extremely high peak, due to Terzaghi stress

reduction only. So at early stage of shale gas production, low Terzaghi stress and highly-

Page 150: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

135

opened fractures significantly enhanced Darcy permeability. Darcy viscous flow is more

critical at this moment though other flow regimes are always happening simultaneously. At

late stage, low pressure and narrow flow path increase the Knudsen number and the

contribution from slip flow and Knudsen diffusion. This is one reason that we need to

integrate different flow models for multiple flow regimes in the apparent permeability to

account for shale gas production.

4.2 Permeability evolution with sorption

In Figure 50, permeability results of He and CO2 gas permeability for two disk

samples (two cycles each) are displayed in the way that we can easily observe and make

comparison. The average magnitude of helium permeability yields a larger value than that

of CO2 permeability when pressure is below 1000 psi. This seems common in sorbing

material gas flow due to the matrix swelling/shrinkage intrigued by gas ad-/desorption (S.

Liu & Harpalani, 2014d). However, at above 1000 psi, the CO2 permeability exceeds

helium permeability and have larger increasing slope. Originally, with adsorption-induced

matrix swelling, permeability is supposed to be reduced due to less pore volume in rock

matrix. Helium permeability should always exceed CO2 permeability for the same sample

following the above hypothesis. So the reason that exactly opposite phenomenon happened

could result from a physical property of shale rocks: its limited gas adsorption capacity.

Low adsorption capacity makes shale sample ceased in adsorbing CO2 molecules at certain

pressure when the maximum adsorption amount is reached. So when the gas pressure is

high enough, the absorption process will stop and sorption-induced deformation on sample

will not contribute to permeability evolution anymore. The unexpected permeability trends

prove that different adsorption behavior for different rocks are very critical in describing

Page 151: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

136

gas permeability evolution. Both sorption-induced deformation and different flow behavior

between gases consist of the difference between helium and CO2 permeability in shale rock.

This can be concluded that we need to combine the influence of sorption-induced rock

deformation and the change of flow regimes to account for shale gas production.

4.3 Permeability evolution with effective stress

In order to further study and address the impact of effective stress on shale permeability,

confining stresses (𝜎𝑐 ) and pore pressures (p) were linearly plotted with logarithm of

apparent permeability (k) to determine the effective stress coefficient (𝜒 ) and the true

effective stress (𝜎𝑒). In Figure 51, log k versus 𝜎𝑐 were plotted at fixed pore pressures for

helium and CO2 injections on 5.39 mm sample. Similar slopes were obtained from isolines

of different pore pressures. Average values were used for effective stress coefficient

calculation, which lead to 𝜕𝑙𝑜𝑔𝑘

𝜕𝜎𝑐= −0.00027 for helium permeability data and

𝜕𝑙𝑜𝑔𝑘

𝜕𝜎𝑐=

−0.000131 for CO2 permeability data. The negative values indicate that apparent

permeabilities decrease with the increase of Terzaghi stresses, which is consistent with

previous finding from other researchers (Heller et al., 2014; Kwon et al., 2001). For 3.91

mm sample, we have -0.000086 and -0.000077 for helium and CO2 data separately. It is

obvious that the slope for helium is always larger than CO2, indicating that helium is more

sensitive to the variation of Terzaghi stress and it is independent of p.

Kwon et al. (2001) and Heller et al. (2014) obtained only one positive slope when

plotting log k versus p for each rock sample. This is probably because when they measured

permeability at low pressure (p<1000 psi), they obtained 𝜒 by fitting high pressure (>1000

psi) permeability data and used the same 𝜒 for the evaluation of low pressure permeabilities.

Page 152: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

137

However, we also want to obtain 𝜒 at low pressure due to two reasons: 1. the apparent

permeabilities we measured have increasing trends when pressure decreased at low

pressure (<500 psi), so the impact of true effective pressure on apparent permeability needs

to evaluate in the same pressure region. 2. If we use the low pressure permeability data to

fit the effective stress law, 𝜒 will likely be changed for same sample. Because the

relationship between logk and p become negative, and it will result in a negative 𝜒 value

eventually. Assuming this ratio-of-slope method is still valid at low pressure (<500 psi),

we plotted two 𝜕𝑙𝑜𝑔𝑘

𝜕𝑝 at constant 𝜎𝑐 for 5.44 mm sample per gas type with breaking down

the pressure at ~500 psi into two divisions. In Figure 52, we can clearly observe that the

slopes for logk versus p are always positive at high pressure division (>500 psi) and

negative at low pressure division (<500 psi). Same behavior happened on 3.91 mm sample.

Averagely, 𝜕𝑙𝑜𝑔𝑘

𝜕𝑝= −0.00039 and

𝜕𝑙𝑜𝑔𝑘

𝜕𝑝= 0.00015 were obtained through plot reading.

The entirely different slopes result from the change of controlling flow regime during

pressure depletion. Knudsen diffusion and slip effect are more intensive at low pressure

due to narrowed flow path, and the permeability is enhanced even effective stress keeps

increasing simultaneously (Heller et al., 2014). Very similar behavior can be observed for

both helium and CO2 injection results, except the absolute values of CO2 slopes are larger

than helium (-0.00051 and 0.00063). This is exactly opposite to the conclusion drawn from

logk versus 𝜎𝑐 data, which indicates CO2 are more sensitive to p change than helium.

Using Eqs. (19), we are able to calculate different 𝜒 values for each individual 𝜕𝑙𝑜𝑔𝑘

𝜕𝑝.

Corresponding to the 𝜕𝑙𝑜𝑔𝑘

𝜕𝑝 values, 𝜒 becomes negative at low pressure (<500 psi) and

Page 153: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

138

positive at high pressure (>500 psi). In Figure 53, 𝜎𝑒 versus k were plotted by fitting

experimental data to the 𝜎𝑒 law for 5.39 mm sample. Although we used two different 𝜒

values on the same permeability curves and the two curves are not located on the same

trajectory, the overall trends of two are quite similar to each other. We can still conclude

that at low pressure region and high pressure region, respectively, the permeability will

decrease with the increase of true effective pressure based on effective stress law, which is

partially consistent with previous finding by Kwon et al. (2001) and Heller et al. (2014).

One of the biggest differences between our work and theirs is that we applied two effective

stress coefficients based on truly observed difference in permeability evolution trends at

different pressure levels. The consistency in permeability trend with two 𝜒 values can

justify the availability of this approach in evaluating permeability evolution with respect

to effective stress. When pressure is larger than 500 psi, 𝜒 was found to be positive and

less than 1, indicating the shale sample is more sensitive to changes in 𝜎𝑐 than changes in

p. Potentially this may result from very low clay content in rock sample (Kwon et al., 2001).

On the other hand, the absolute value of 𝜒 is larger than 1 at low pressure (<500 psi),

meaning permeability is more sensitive to pore pressure change. The discrepancy between

𝜒 values for same rock sample same gas results from the difference of which controlling

factor, 𝜎𝑐 or p, dominates gas flow in shale pores. Compared with helium 𝜒 values for 5.44

mm sample, CO2 results are 3 to 6 times larger at either low or high pressure region, which

seems to gives us the information that CO2 permeability can be extremely sensitive to the

change of p rather than 𝜎𝑐. However, the 𝜒 values of 3.91 mm sample are -5.226, 1.151

and -1.693, 1.324 for helium and CO2, respectively. There are no obvious patterns based

Page 154: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

139

on existing 𝜒 data. Another difference between helium and CO2 results in Figure 53 is that

the modified permeability curves in high pressure region (>500 psi) accompany with

negative true effective stresses. It happens when 𝜒 values are extremely high that the

proportional contribution of p on 𝜎𝑒 is overestimated. For 3.91 mm, the modified

permeability curves will locate in positive stress region since the 𝜒 values are not deviated

and close to one.

4.4 Modeling results of Knudsen diffusion flow

From the above permeability data profiles and analysis, we can see how multiple flow

regimes control gas permeability within different pressure region, the significance of

effective stress and the sorption-induced matrix/pore deformation on permeability

evolution. With different type of flow regimes, different external stresses and different

gases, permeability on shale can show totally diverse trends and values during pulse-decay

permeability measurements and even in real production process. Since late-time production

tail can be quite sustainable and Knudsen diffusion is dominant at low pressure in this

period, we firstly want to conduct modeling study for Knudsen diffusion only to figure out

how it may contribute to the total shale gas permeability.

Figure 54 gives results on the Knudsen apparent permeability change at our lab

condition for both helium and CO2. A range of pore size from 5-500 nm was assigned to

cover possible flow pathways inside the shale disk sample. The permeability evolution

trends shown in these two figures are quite similar to each other. Knudsen gas apparent

permeability values are inversely proportional to gas pressure. The reduction of from very

low pressure to high pressure can be extremely large, explaining that the Knudsen diffusion

Page 155: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

140

at high pressure really plays no roles. With an increasing average pore size, the overall

permeability becomes larger due to enlarged flow path. Compared to average helium

apparent diffusive permeability, average CO2 apparent diffusive permeability is more than

three times lower. Purely excluding the adsorption effect, this result indicates that helium

molecules have more intensive and easier diffusion flow process than CO2 molecules do in

the same medium. On the other hand, the apparent diffusive permeability also increases

significantly with increasing average pore size. This behavior is similar to Darcy flow, but

it may not be valid when pore size is large enough when there are more flow regimes

simultaneously happening.

For the purpose of permeability modeling, the permeability values for helium and CO2

are converted to ratios between permeability and initial permeability, shown in Figure 55

and 56. Instead of different declining curves of Knudsen apparent permeability values,

only a single representative permeability ratio trend can be obtained regardless of pore size

variation. This indicates that the Knudsen apparent permeability can be quite different in

values but will very likely be the same in trend. However, when compared to the laboratory

data obtained in this study, the Knudsen permeability modeling curve displayed

mismatches. Although this declining trend was reported by several laboratory and

modeling works on unconventional rock permeability (Javadpour, 2009; J. Li, Liu, Lu, Yao,

& Xue, 2015; Sakhaee-Pour & Bryant, 2012; Ji Quan Shi & Durucan, 2014; Singh &

Javadpour, 2015), the significant contribution of effective stress and sorption-induced

deformation on permeability should not be ignored (Heller et al., 2014; S. Liu & Harpalani,

2013b, 2014e). In this case, we can find that the initial decreasing trend of Knudsen

apparent permeability more or less can represent what happens on experimental data at low

Page 156: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

141

pressure. However, it lacks the turning up point that can be described by the reduction of

effective stress when pressure keeps increasing. So for both result figures the permeability

data are more than 10 times larger than the Knudsen apparent permeability and have the

potential to keep increasing at higher pressure. Obvious discrepancy is exposed on both

helium and CO2 results. At this point, we may see that only Knudsen diffusion flow model

cannot give a good explanation on shale permeability evolution in a whole picture.

Knudsen diffusion is definitely critical in nano-tube granular flow investigation and

dominating at late-time shale gas production process. However, we still need to combine

multiple flow regimes, effective stress and sorption-induced pore deformation effect to

construct a deliberate study on shale gas flow characterization.

5. Conclusions

A laboratory study has been conducted to investigate the gas (He, CO2) transport in

shale thin disks (matrix flow). Discussion among different flow regimes mainly including

Darcy viscous flow and Knudsen diffusion were carried out. Knudsen diffusion flow, as a

critical flow regime especially in late-time shale gas production, was addresses with

laboratory and modeling results. Based on the work completed, the following conclusions

are made and summarized:

1. The main flow regimes are dynamically changing during shale gas production, and all

flow regimes including Darcy viscous flow, slip flow, transition flow and Knudsen

diffusion exist simultaneous with only difference in their proportional contribution.

2. The non-Darcy flow effect, especially Knudsen flow in this study, has significant

influence on shale gas behavior at low pressure/late-time production.

3. Helium permeability is more sensitive to the variation of Terzaghi stress than CO2 and

Page 157: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

142

it is independent of pore pressure. Apparent permeability decreases when Terzaghi

stress increases.

4. The relationship between apparent permeability and pore pressure can result in both

positive and negative slopes. The negative slope indicates the Knudsen diffusion effect

and slip flow effect have more significant enhancement on permeability than the

reduction from increasing Terzaghi stress. So the true effective stress coefficient may

be calculated separately within different pressure regions < ~500 psi and > ~500 psi.

5. CO2 apparent permeability can be higher than He at pressure higher than 1000 psi,

which may result from limited shale adsorption capacity, and thus limited sorption-

induced matrix swelling effect when pressure keeps increasing. However, on the

perspective of the analysis based on effective stress law didn’t provide obvious

evidence on the difference between helium and CO2 permeability evolution.

6. Knudsen diffusive permeability can be significantly varied along with average pore

size of shale. Helium molecules can have more intensive and easier diffusion flow

process than CO2 molecules do.

7. Knudsen diffusion is critical in determine late-time shale gas production, but it is far

from successfully describing the full view of shale gas flow mechanism. The effects of

different flow regimes, effective stress and sorption-induced matrix/pore deformation

are badly required in future shale reservoir characterization and shale gas

permeability/flow modeling.

Page 158: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

143

References

Aminian, K., & Ameri, S. (2004). Type Curves for Coalbed Methane Production

Prediction. 2004 SPE Eastern Regional Meeting. http://doi.org/10.2523/91482-MS

Amyx, J. W., Bass, D. M. J., & Whiting, R. L. (1960). Petroleum Reservoir Engineering.

An, F. H., Cheng, Y. P., Wu, D. M., & Wang, L. (2013). The effect of small micropores

on methane adsorption of coals from Northern China. Adsorption, 19(1), 83–90.

http://doi.org/10.1007/s10450-012-9421-3

Brace, W. F., Walsh, J. B., & Frangos, W. T. (1968). Permeability of granite under high

pressure. Journal of Geophysical Research, 73(6). Retrieved from

http://onlinelibrary.wiley.com/doi/10.1029/JB073i006p02225/full

Breckels, I. M., & Eekelen, H. a. M. (1982). Relationship Between Horizontal Stress and

Depth in Sedimentary Basins. Journal of Petroleum Technology, 34(9), 2191–2199.

http://doi.org/10.2118/10336-PA

Busch, A., Gensterblum, Y., Krooss, B. M., & Littke, R. (2004). Methane and carbon

dioxide adsorption–diffusion experiments on coal: upscaling and modeling.

International Journal of Coal Geology, 60(2–4), 151–168.

http://doi.org/10.1016/j.coal.2004.05.002

Cai, Y., Liu, D., Pan, Z., Yao, Y., Li, J., & Qiu, Y. (2013). Pore structure and its impact

on CH4 adsorption capacity and flow capability of bituminous and subbituminous

coals from Northeast China. Fuel, 103, 258–268.

http://doi.org/10.1016/j.fuel.2012.06.055

Choi, J.-G., Do, D. D., & Do, H. D. (2001). Surface Diffusion of Adsorbed Molecules in

Porous Media: Monolayer, Multilayer, and Capillary Condensation Regimes.

Industrial & Engineering Chemistry Research, 40(19), 4005–4031.

http://doi.org/10.1021/ie010195z

Civan, F., Rai, C., & Sondergeld, C. (2011). Shale-gas permeability and diffusivity

inferred by improved formulation of relevant retention and transport mechanisms.

Transport in Porous Media, 86(3), 925–944. http://doi.org/10.1007/s11242-010-

9665-x

Clarkson, C., & Bustin, R. (1999). 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, di, 1333–1344. Retrieved from

http://www.sciencedirect.com/science/article/pii/S0016236199000551

Clarkson, C., Pan, Z., Palmer, I., & Harpalani, S. (2010). Predicting sorption-induced

strain and permeability increase with depletion for coalbed-methane reservoirs. SPE

Journal, (March).

Clarkson, C. R., & Bustin, R. M. (1999). The effect of pore structure and gas pressure

upon the transport properties of coal: a laboratory and modeling study. 2. Adsorption

rate modeling. Fuel, 78(11), 1345–1362. http://doi.org/10.1016/S0016-

Page 159: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

144

2361(99)00056-3

Clarkson, C. R., Solano, N., & Bustin, R. (2013). Pore structure characterization of North

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

intrusion. Fuel, 103, 606–616. http://doi.org/10.1016/j.fuel.2012.06.119

Coal Age - DEC 2014. (2014).

Coppens, M. (1999). The effect of fractal surface roughness on diffusion and reaction in

porous catalysts – from fundamentals to practical applications. Catalysis Today,

53(2), 225–243. http://doi.org/10.1016/S0920-5861(99)00118-2

Coppens, M. O., & Dammers, A. J. (2006). Effects of heterogeneity on diffusion in

nanopores - From inorganic materials to protein crystals and ion channels. Fluid

Phase Equilibria, 241(1–2), 308–316. http://doi.org/10.1016/j.fluid.2005.12.039

Crank, J. (1957). The mathematics of diffusion.

Cui, X., Bustin, A., & Bustin, R. (2009). Measurements of gas permeability and

diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 40479.

Cui, X., Bustin, A. M. M., & Bustin, R. M. (2009). Measurements of gas permeability

and diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 9(3), 208–223. http://doi.org/10.1111/j.1468-8123.2009.00244.x

Cui, X., & Bustin, R. M. (2005). Volumetric strain associated with methane desorption

and its impact on coalbed gas production from deep coal seams. AAPG Bulletin,

89(9), 1181–1202. http://doi.org/10.1306/05110504114

Cui, X., Bustin, R. M., & Dipple, G. (2004). Selective transport of CO2, CH4, and N2 in

coals: insights from modeling of experimental gas adsorption data. Fuel, 83(3), 293–

303. http://doi.org/10.1016/j.fuel.2003.09.001

Darabi, H., Ettehad, a., Javadpour, F., & Sepehrnoori, K. (2012). Gas flow in ultra-tight

shale strata. Journal of Fluid Mechanics, 710, 641–658.

http://doi.org/10.1017/jfm.2012.424

Dicker, A. I., & Smits, R. M. (1988). A practical approach for determining permeability

from laboratory pressure-pulse decay measurements. In SPE International Meeting

on Petroleum Engineering, Tianjin, China,1-4 Nov,1998. Tianjin, China. Retrieved

from https://www.onepetro.org/conference-paper/SPE-17578-MS

Do, D. D. (1998). Adsorption Analysis: Equilibria and Kinetics (Vol. 2). Imperial

College Press. http://doi.org/10.1142/9781860943829

Drazer, G., Chertcoff, R., Bruno, L., & Rosen, M. (1999). Concentration dependence of

diffusion–adsorption rate in activated carbon. Chemical Engineering Science, 54,

4285–4291.

Dutta, A. (2009). Multicomponent Gas Diffusion and Adsorption in Coals for Enhanced

Methane Recovery.

EIA. (2015). Marcellus Region Drilling Productivity Report, (March), 2015.

Page 160: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

145

Ertekin, T., King, G., & Schwerer, F. (1986). Dynamic Gas Slippage: A Unique Dual-

Mechanism Approach to the Flow of Gas in Tight Formations. SPE Formation

Evaluation. http://doi.org/10.2118/12045-PA

Gan, H., Nandi, S. P., & Walker, P. L. (1972). Nature of the porosity in American coals.

Fuel, 51(4), 272–277. http://doi.org/10.1016/0016-2361(72)90003-8

Geertsma, J. (1966). Problems of rock mechanics in petroleum production engineering.

1st ISRM Congress, 3(59), 585–594.

Gensterblum, Y., Ghanizadeh, A., & Krooss, B. M. (2014). Gas permeability

measurements on Australian subbituminous coals: Fluid dynamic and poroelastic

aspects. Journal of Natural Gas Science and Engineering, 19, 202–214.

http://doi.org/10.1016/j.jngse.2014.04.016

Halliburton Company. (2007). Chapter 04 - Reservoir Analysis. Coalbed Methane

Princples and Practices, (June 2007), 191–282.

Harpalani, S., & Chen, G. (1997). Influence of gas production induced volumetric strain

on permeability of coal. Geotechnical and Geological Engineering, 15(4), 303–325.

http://doi.org/10.1007/BF00880711

Harpalani, S., & Mitra, A. (2009). Impact of CO2 Injection on Flow Behavior of Coalbed

Methane Reservoirs. Transport in Porous Media, 82(1), 141–156.

http://doi.org/10.1007/s11242-009-9475-1

Harpalani, S., Prusty, B., & Dutta, P. (2006). Methane/CO2 sorption modeling for

coalbed methane production and CO2 sequestration. Energy & Fuels, 20(4), 1591–

1599. Retrieved from http://pubs.acs.org/doi/abs/10.1021/ef050434l

Hartman, R. (2008). Recent advances in the analytical methods used for shale gas

reservoir gas-in-place assessment. In AAPG Annual Convention, San Antonio,

Texas, 20-23 Apr, 2008 (Vol. 40317).

He, W., Lv, W., & Dickerson, J. (2014). Gas Diffusion Mechanisms and Models. In Gas

Transport in Solid Oxide Fuel Cells (pp. 9–17). http://doi.org/10.1007/978-3-319-

09737-4

Heller, R., Vermylen, J., & Zoback, M. (2014). Experimental investigation of matrix

permeability of gas shales. AAPG Bulletin, 98(5), 975–995.

http://doi.org/10.1306/09231313023

Hildenbrand, a., Krooss, B. M., Busch, a., & Gaschnitz, R. (2006). Evolution of

methane sorption capacity of coal seams as a function of burial history — a case

study from the Campine Basin, NE Belgium. International Journal of Coal Geology,

66(3), 179–203. http://doi.org/10.1016/j.coal.2005.07.006

Hsieh, P. a., Tracy, J. V., Neuzil, C. E., Bredehoeft, J. D., & Silliman, S. E. (1981). A

transient laboratory method for determining the hydraulic properties of “tight”

rocks—I. Theory. International Journal of Rock Mechanics and Mining Sciences &

Geomechanics Abstracts, 18(3), 245–252. http://doi.org/10.1016/0148-

9062(81)90979-7

Page 161: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

146

Hubbert, M. K. (1957). Darcy’S Law and the Field Equations of the Flow of

Underground Fluids. International Association of Scientific Hydrology. Bulletin,

2(1), 23–59. http://doi.org/10.1080/02626665709493062

Izadi, G., & Elsworth, D. (2013). The effects of thermal stress and fluid pressure on

induced seismicity during stimulation to production within fractured reservoirs.

Terra Nova, 25(5), 374–380. http://doi.org/10.1111/ter.12046

Izadi, G., Wang, S., Elsworth, D., Liu, J., Wu, Y., & Pone, D. (2011). Permeability

evolution of fluid-infiltrated coal containing discrete fractures. International Journal

of Coal Geology, 85(2), 202–211. http://doi.org/10.1016/j.coal.2010.10.006

Javadpour, F. (2009). Nanopores and apparent permeability of gas flow in mudrocks

(shales and siltstone). Journal of Canadian Petroleum Technology, 48, 1–6.

Javadpour, F., Fisher, D., & Unsworth, M. (2007). Nanoscale gas flow in shale gas

sediments. Journal of Canadian Petroleum Technology, 46(10).

http://doi.org/10.2118/07-10-06

Jones, S. (1997). A technique for faster pulse-decay permeability measurements in tight

rocks. SPE Formation Evaluation, (November 1994), 25–28.

Kamath, J., Boyer, R., & Nakagawa, F. (1992). Characterization of core-scale

heterogeneities using laboratory pressure transients. SPE Formation Evaluation,

(September), 219–227.

Karniadakis, G., Beskok, A., & Narayan, A. (2005). Microflows and Nanoflows.

http://doi.org/10.1007/0-387-28676-4

Klinkenberg, L. J. (1941). The permeability of porous media to liquids and gases.

Drilling and Production Practice, (2), 57–73.

http://doi.org/10.5510/OGP20120200114

Kuila, U., Prasad, M., & Kazemi, H. (2013). Assessing Knudsen flow in gas-flow models

of shale reservoirs (pp. 23–28).

Kumar, A. (2007). Methane Diffusion Characteristics of Illinois Coals.

Kumar, H., Elsworth, D., Liu, J., Pone, D., & Mathews, J. P. (2012). Optimizing

enhanced coalbed methane recovery for unhindered production and CO2 injectivity.

International Journal of Greenhouse Gas Control, 11, 86–97.

http://doi.org/10.1016/j.ijggc.2012.07.028

Kumar, H., Elsworth, D., Mathews, J. P., Liu, J., & Pone, D. (2014). Effect of CO2

injection on heterogeneously permeable coalbed reservoirs. Fuel, 135, 509–521.

http://doi.org/10.1016/j.fuel.2014.07.002

Kumar, H., Elsworth, D., Mathews, J. P., & Marone, C. (2015). Permeability evolution in

sorbing media: analogies between organic-rich shale and coal. Geofluids.

http://doi.org/10.1111/gfl.12135

Kwon, O., Kronenberg, A. K., Gangi, A. F., & Johnson, B. (2001). Permeability of

Wilcox shale and its effective pressure law. Journal of Geophysical Research, 106,

19339. http://doi.org/10.1029/2001JB000273

Page 162: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

147

Langmuir, I. (1918). The adsorption of gases on plane surfaces of glass, mica and

platinum. Journal of the American Chemical Society, 345(1914). Retrieved from

http://pubs.acs.org/doi/pdf/10.1021/ja02242a004

Levine, J. (1996). Model study of the influence of matrix shrinkage on absolute

permeability of coal bed reservoirs. Geological Society, London, Special

Publications.

Li, J., Liu, D., Lu, S., Yao, Y., & Xue, H. (2015). Evaluation and modeling of the CO2

permeability variation by coupling effective pore size evolution in anthracite coal.

Energy & Fuels. http://doi.org/10.1021/ef502674y

Li, J., Liu, D., Yao, Y., Cai, Y., & Chen, Y. (2013). Evaluation and modeling of gas

permeability changes in anthracite coals. Fuel, 111, 606–612.

http://doi.org/10.1016/j.fuel.2013.03.063

Li, Y., Li, X., Shi, J., Wang, H., Wu, L., Teng, S., & Corp, S. (2014). A Nano-Pore Scale

Gas Flow Model for Shale Gas Reservoir Flow characteristic of shale gas reservoir

in various pole scales and molecule scales, (June), 9–11.

Liu, J., Chen, Z., Elsworth, D., Miao, X., & Mao, X. (2010). Linking gas-sorption

induced changes in coal permeability to directional strains through a modulus

reduction ratio. International Journal of Coal Geology, 83(1), 21–30.

http://doi.org/10.1016/j.coal.2010.04.006

Liu, S. (2012). Estimation of different coal compressibilities of coalbed methane

reservoirs under replicated in situ condition.

Liu, S., & Harpalani, S. (2013a). A new theoretical approach to model sorption-induced

coal shrinkage or swelling. AAPG Bulletin, 97(7), 1033–1049.

http://doi.org/10.1306/12181212061

Liu, S., & Harpalani, S. (2013b). Permeability prediction of coalbed methane reservoirs

during primary depletion. International Journal of Coal Geology, 113, 1–10.

http://doi.org/10.1016/j.coal.2013.03.010

Liu, S., & Harpalani, S. (2014a). Compressibility of sorptive porous media: Part 1.

Background and theory. AAPG Bulletin, c, 1–25.

Liu, S., & Harpalani, S. (2014b). Compressibility of Sorptive Porous Media: Part I –

Background and Theory. AAPG Bulletin, 98(9), 1761–1772.

Liu, S., & Harpalani, S. (2014c). Compressibility of Sorptive Porous Media: Part II –

Experimental Study on Coal. AAPG Bulletin, 98(9), 1773–1788.

Liu, S., & Harpalani, S. (2014d). Determination of the effective stress law for

deformation in coalbed methane reservoirs. Rock Mechanics and Rock Engineering,

47(5), 1809–1820. http://doi.org/10.1007/s00603-013-0492-6

Liu, S., & Harpalani, S. (2014e). Evaluation of in situ stress changes with gas depletion

of coalbed methane reservoirs. Journal of Geophysical Research: Solid Earth,

119(8), 6263–6276. http://doi.org/10.1002/2014JB011228.Received

Liu, S., Harpalani, S., & Pillalamarry, M. (2012). Laboratory measurement and modeling

Page 163: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

148

of coal permeability with continued methane production: Part 2 – Modeling results.

Fuel, 94, 117–124. http://doi.org/10.1016/j.fuel.2011.10.053

Lorenz, J. C., Teufel, L. W., & Warpinski, N. R. (1991). Regional fractures; I, A

mechanism for the formation of regional fractures at depth in flat-lying reservoirs.

AAPG Bulletin , 75(11), 1714–1737. Retrieved from

http://aapgbull.geoscienceworld.org/content/75/11/1714.abstract

Luffel, D., Hopkins, C., & Jr, P. S. (1993). Matrix permeability measurement of gas

productive shales. In SPE Annual Technical Conference and Exhibition, Houston,

Texas, 2-6 Oct, 1993.

Ma, Q., Harpalani, S., & Liu, S. (2011). A simplified permeability model for coalbed

methane reservoirs based on matchstick strain and constant volume theory.

International Journal of Coal Geology, 85(1), 43–48.

http://doi.org/10.1016/j.coal.2010.09.007

Malkovsky, V. I., Zharikov, a. V., & Shmonov, V. M. (2009). New methods for

measuring the permeability of rock samples for a single-phase fluid. Izvestiya,

Physics of the Solid Earth, 45(2), 89–100.

http://doi.org/10.1134/S1069351309020013

Markowski, A. K. (2014). PENNSYLVANIA COALBED METHANE UPDATE, 80353.

Markowski, A., Resources, N., & Virginia, W. (2014). Coalbed methane in

Pennsylvania.

Mavor, M. J., Owen, L. B., & Pratt, T. J. (1990). Measurement and Evaluation of Coal

Sorption Isotherm Data. Proceedings of SPE Annual Technical Conference and

Exhibition, (1). http://doi.org/10.2523/20728-MS

Mazumder, S., & Wolf, K. (2008). Differential swelling and permeability change of coal

in response to CO 2 injection for ECBM. International Journal of Coal Geology.

Mckernan, R. E., Rutter, E. H., Mecklenburgh, J., & Taylor, K. G. (2014). Influence of

Effective Pressure on Mudstone Matrix Permeability: Implications for Shale Gas

Production. In SPE/EAGE European Unconventional Conference and Exhibition,

Vienna, Austria, 25-27 Feb, 2014 (p. 13). Society of Petroleum Engineers.

http://doi.org/10.2118/167762-MS

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1–Laboratory results. Fuel,

94, 110–116.

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1 – Laboratory results. Fuel,

94, 110–116. http://doi.org/10.1016/j.fuel.2011.10.052

Okotie, V. U., & Moore, R. L. (2011). Well-production challenges and solutions in a

mature, very-low-pressure coalbed-methane reservoir. SPE Production and

Operations, 26(2), 149–161. http://doi.org/10.2118/107705-MS;

Palmer, I., & Mansoori, J. (1998). How permeability depends on stress and pore pressure

Page 164: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

149

in coalbeds: a new model. SPE Reservoir Evaluation and Engineering, SPE

52607(December), 539–544.

Palmer, S. I., & Mansoori, J. (1998). How permeability depends on stress and pore

pressure in coalbeds: a new model. Soc Petrol Eng Res Eng, (December), 539–544.

Retrieved from

http://www.pe.tamu.edu/wattenbarger/public_html/Selected_papers/--Coalbed

Methane/SPE 36737 Palmer and Mansoori.pdf

Pan, Z., & Connell, L. D. (2007). A theoretical model for gas adsorption-induced coal

swelling. International Journal of Coal Geology, 69(4), 243–252.

http://doi.org/10.1016/j.coal.2006.04.006

Pillalamarry, M., Harpalani, S., & Liu, S. (2011). Gas diffusion behavior of coal and its

impact on production from coalbed methane reservoirs. International Journal of

Coal Geology, 86(4), 342–348. http://doi.org/10.1016/j.coal.2011.03.007

Randolph, P. L., Soeder.D.J., & Chowdiah, P. (1984). Porosity and Permeability of Tight

Sands. In SPE/DOE/GRI Unconventional Gas Recovery Symposium held in

Pittsburgh, PA, May 13-15. Dallas.

Reif.F. (1965). Fundamentals of statistical and thermal physics. Retrieved from

http://catalog.hathitrust.org/Record/000145539/Cite

Rexer, T. F. T., Benham, M. J., Aplin, A. C., & Thomas, K. M. (2013). Methane

Adsorption on Shale under Simulated Geological Temperature and Pressure

Conditions. Energy & Fuels, 27(6), 3099–3109. http://doi.org/10.1021/ef400381v

Rodrigues, C. F., & Lemos De Sousa, M. J. (2002). The measurement of coal porosity

with different gases. International Journal of Coal Geology, 48(3–4), 245–251.

http://doi.org/10.1016/S0166-5162(01)00061-1

Roy, S., Raju, R., Chuang, H. F., Cruden, B. a., & Meyyappan, M. (2003). Modeling gas

flow through microchannels and nanopores. Journal of Applied Physics, 93(8),

4870–4879. http://doi.org/10.1063/1.1559936

Ruckenstein, E. (1971). Sorption by solids with bidisperse pore structures. Chemical

Engineering …, 26, 1305–1318.

Rydzy, M. B., Patino, J., Elmetni, N., & Appel, M. (2016). STRESSED

PERMEABILITY IN SHALES : EFFECTS OF MATRIX COMPRESSIBILITY

AND FRACTURES - A STEP TOWARDS MEASURING MATRIX

PERMEABILITY IN FRACTURED SHALE SAMPLES, 1–12.

Sakhaee-Pour, A., & Bryant, S. (2012). Gas Permeability of Shale. SPE Reservoir

Evaluation & Engineering, 15, 401–409.

Seidle, J. P., Jeansonne, M. W., & Erickson, D. J. (1992). Application of matchstick

geometry to stress dependent permeability in coals. SPE Rocky Mountain Regional

Meeting, 433–444. http://doi.org/10.2523/24361-MS

Shabro, V. (2012). Forecasting Gas Production in Organic Shale with the Combined

Numerical Simulation of Gas Diffusion in Kerogen, Langmuir Desorption from

Page 165: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

150

Kerogen Surfaces, and Advection in Nanopores. SPE Annual Technical ….

http://doi.org/10.2118/159250-MS

Shi, J.-Q., Pan, Z., & Durucan, S. (2014). Analytical models for coal permeability

changes during coalbed methane recovery: Model comparison and performance

evaluation. International Journal of Coal Geology, 136, 17–24.

http://doi.org/10.1016/j.coal.2014.10.004

Shi, J., & Durucan, S. (2005). A model for changes in coalbed permeability during

primary and enhanced methane recovery. SPE Reservoir Evaluation & Engineering,

8(4), 291–299.

Shi, J. Q., & Durucan, S. (2003). A bidisperse pore diffusion model for methane

displacement desorption in coal by CO 2 injection☆. Fuel, 82(10), 1219–1229.

http://doi.org/10.1016/S0016-2361(03)00010-3

Shi, J. Q., & Durucan, S. (2004). Drawdown Induced Changes in Permeability of

Coalbeds: A New Interpretation of the Reservoir Response to Primary Recovery.

Transport in Porous Media, 56(1), 1–16.

http://doi.org/10.1023/B:TIPM.0000018398.19928.5a

Shi, J. Q., & Durucan, S. (2014). Modelling laboratory horizontal stress and coal

permeability data using S&D permeability model. International Journal of Coal

Geology, 131, 172–176. http://doi.org/10.1016/j.coal.2014.06.014

Singh, H., & Javadpour, F. (2013). Nonempirical Apparent Permeability of Shale. In

Unconventional Resources Technology Conference (Vol. 2, p. 16). Denvor,

Colorado.

Singh, H., & Javadpour, F. (2015). Langmuir slip-Langmuir sorption permeability model

of shale. FUEL, (October). http://doi.org/10.1016/j.fuel.2015.09.073

Smith, D., & Williams, F. (1984). Diffusional effects in the recovery of methane from

coalbeds. SPEJ, Soc. Pet. Eng. J.;(United States), (October), 529–535.

Soeder, D. (1988). Porosity and permeability of eastern Devonian gas shale. SPE

Formation Evaluation, 3, 116–124.

Sudibandriyo, M., Pan, Z., Fitzgerald, J. E., Jr, R. L. R., & Gasem, K. A. M. (2003).

Adsorption of Methane, Nitrogen, Carbon Dioxide, and Their Binary Mixtures on

Dry Activated Carbon at 318.2 K and Pressures up to 13.6 MPa. Lan, 19(8), 5323–

5331.

Suggested Methods for Uniaxial‐Strain Compressibility Testing for Reservoir

Geomechanics. (2015), c(April), 1–29.

Swami, V. (2012). Shale Gas Reservoir Modeling: From Nanopores to Laboratory. SPE

Annual Technical Conference and Exhibition, (October), 8–10.

Swami, V., Clarkson, C., & Settari, A. (2012). Non-Darcy Flow in Shale Nanopores: Do

We Have a Final Answer? SPE Canadian Unconventional …, (November), 1–17.

Swami, V., & Settari, A. (2012). A pore scale gas flow model for shale gas reservoir. In

Page 166: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

151

SPE Americas Unconventional Resources Conference held in Pittsburgh (pp. 5–7).

Terzaghi, K. (1943). Theoretical Soil Mechanics. John Wiley & Sons, Inc.

http://doi.org/10.1680/geot.1964.14.1.1

Teufel, L. W., Rhett, D. W., & Farrell, H. E. (1991). Effect of Reservoir Depletion And

Pore Pressure Drawdown On In Situ Stress And Deformation In the Ekofisk Field,

North Sea. The 32nd U.S. Symposium on Rock Mechanics (USRMS), 63–72.

U S Energy Information Administration. (2011). Emissions of Greenhouse Gases in the

United States 2009, (March).

Walls, J., Diaz, E., & Cavanaugh, T. (2012). Shale Reservoir Properties from Digital

Rock Physics. Proceedings of SPE/EAGE European Unconventional Resources

Conference and Exhibition, 1–9. http://doi.org/10.2118/152752-MS

Wang, G., Ren, T., Wang, K., & Zhou, A. (2014). Improved apparent permeability

models of gas flow in coal with Klinkenberg effect. Fuel, 128, 53–61.

http://doi.org/10.1016/j.fuel.2014.02.066

Wang, S., Elsworth, D., & Liu, J. (2011). Permeability evolution in fractured coal: the

roles of fracture geometry and water-content. International Journal of Coal

Geology, 87(1), 13–25. http://doi.org/10.1016/j.coal.2011.04.009

Wang, S., Elsworth, D., & Liu, J. (2012). A mechanistic model for permeability evolution

in fractured sorbing media. Journal of Geophysical Research, 117(B6).

Wang, S., Elsworth, D., & Liu, J. (2012a). A mechanistic model for permeability

evolution in fractured sorbing media. Journal of Geophysical Research: Solid Earth

( …, 117(B6), B06205. http://doi.org/10.1029/2011JB008855

Wang, S., Elsworth, D., & Liu, J. (2012b). Permeability evolution during progressive

deformation of intact coal and implications for instability in underground coal

seams. International Journal of Rock Mechanics and Mining Sciences, 1–12.

Wang, Y., & Liu, S. (2016). Estimation of Pressure-Dependent Diffusive Permeability of

Coal Using Methane Diffusion Coefficient: Laboratory Measurements and

Modeling. Energy and Fuels, 30(11), 8968–8976.

http://doi.org/10.1021/acs.energyfuels.6b01480

Wang, Y., Liu, S., & Elsworth, D. (2015). Laboratory investigations of gas flow

behaviors in tight anthracite and evaluation of different pulse-decay methods on

permeability estimation. International Journal of Coal Geology, 149, 118–128.

http://doi.org/10.1016/j.coal.2015.07.009

Wu, K., Li, X., Wang, C., Chen, Z., & Yu, W. (2015). A Model for Gas Transport in

Microfractures of Shale and Tight Gas Reservoirs. AIChE Journal, 61(6).

http://doi.org/10.1002/aic

Wu, K., Li, X., Wang, C., Yu, W., & Chen, Z. (2014). Apparent Permeability for Gas

Flow in Shale Reservoirs Coupling Effects of Gas Diffusion and Desorption.

Proceedings of the 2nd Unconventional Resources Technology Conference, 1–18.

http://doi.org/10.15530/urtec-2014-1921039

Page 167: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

152

Xu, H., Tang, D., Zhao, J., Li, S., & Tao, S. (2015). A new laboratory method for

accurate measurement of the methane diffusion coefficient and its influencing

factors in the coal matrix. Fuel, 158, 239–247.

http://doi.org/10.1016/j.fuel.2015.05.046

Yi, J., Akkutlu, I., & Deutsch, C. (2008). Gas Transport in Bidisperse Coal Particles:

Investigation for an Effective Diffusion Coefficient in Coalbeds. Journal of

Canadian Petroleum Technology, 47(10). http://doi.org/10.2118/08-10-20

Yin, G., Jiang, C., Wang, J. G., & Xu, J. (2013). Combined Effect of Stress, Pore

Pressure and Temperature on Methane Permeability in Anthracite Coal: An

Experimental Study. Transport in Porous Media, 100(1), 1–16.

http://doi.org/10.1007/s11242-013-0202-6

Yuan, W., Pan, Z., Li, X., Yang, Y., Zhao, C., Connell, L. D., … He, J. (2014).

Experimental study and modelling of methane adsorption and diffusion in shale.

Fuel, 117, 509–519. http://doi.org/10.1016/j.fuel.2013.09.046

Zhang, T., Ellis, G. S., Ruppel, S. C., Milliken, K., & Yang, R. (2012). Effect of organic-

matter type and thermal maturity on methane adsorption in shale-gas systems.

Organic Geochemistry, 47, 120–131.

http://doi.org/10.1016/j.orggeochem.2012.03.012

Zoback, M. D., & Byerlee, J. D. (1975). Permeability and Effective Stress. Aapg Bulletin-

American Association of Petroleum Geologists, 59(1), 154–158.

http://doi.org/10.1306/83D91C40-16C7-11D7-8645000102C1865D

Page 168: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

153

Appendix

Figure 43 Average monthly gas production of 15 production wells in Eagleford Shale Group with

flow regimes breakdown from large-scale Darcy/continuum flow to Knudsen diffusion/molecular

flow.

Darcy/ Continuum

Flow Slip Flow Transition

Flow

Knudsen/Free

Molecule Flow

0 0.001 0.1 1 10 𝑲𝒏

Pressure-driven

Concentration-driven and

Molecular Dynamics

Darcy + Knudsen

Page 169: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

154

Figure 44 Gas desorption process in nano-scale. During gas production, gas molecules firstly

desorb from the kerogen surface to the pore, and then move outward due to pressure/concentration

drop. Then kerogen tend to shrink, together with the matrix.

Figure 45 𝐾𝑛 changing with pressure and average pore size for CO2 at 296 K. Mostly the flow is

lying in the transition flow and slip flow region.

Page 170: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

155

Figure 46 Photograph of a slice of Marcellus shale drilled core sample cut into a disk with 5.39 mm

in thickness and 1 inch in diameter.

Figure 47 Schematic view of the pulse-decay experimental setup for permeability evolution test.

V1 is the valve controlling inlet gas flow from upstream, V2 is the valve controlling outlet gas flow

to downstream and V3/V4 is the valve controlling the confining/axial stress applied on the sample.

Page 171: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

156

Figure 48 Pulse-decay experimental data with He and CO2 injection on shale thin disks (5.39 mm

& 3.91 mm) under 1600 psi confining (hydrostatic) stress, processed by using Cui et al.’s method.

Page 172: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

157

Figure 49 Permeability data with helium and CO2 injection on shale thin disks (5.39 mm & 3.91

mm) under 3000 psi confining (hydrostatic) stress, processed by using Cui et al.’s method.

Page 173: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

158

Figure 50 Comparison between averaged helium and CO2 permeability data of 5.39 mm sample.

Figure 51 Helium and CO2 permeability data of 5.39 mm shale disc linearized by plotting logarithm

of permeability versus confining stresses at constant pore pressures to determine the sensitivity of

confining stress to permeability.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 500 1000 1500 2000 2500

Per

mea

bil

ity (

mD

)

Pore Pressure (psi)

He 1600psi Avg

CO2 1600psi Avg

He 3000psi Avg

CO2 3000psi Avg

Page 174: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

159

Figure 52 Helium and CO2 permeability data of 5.39 mm shale disc linearized by plotting logarithm

of permeability versus p at constant confining stresses to determine the sensitivity of pore pressure

to permeability. The slopes are positive when pressures are above ~500 psi and negative when

below ~500 psi.

Page 175: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

160

Figure 53 Permeability versus modified effective stress by fitting experimental data to the effective

stress law for 5.39 mm sample. There are two effective stress coefficients per sample per gas

because of the separated slopes for the permeability-pore pressure relationships.

Page 176: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

161

Figure 54 Permeability values of Knudsen flow permeability model within the same range of

average pore size (5~500 nm) for helium and CO2.

Figure 55 Shale helium injection data and Knudsen diffusion permeability model comparison.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

50 500 5000

k/k

0

Gas Pressure (psi)

Modeling

3.91mm data

5.44mm data

Page 177: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

162

Figure 56 Shale CO2 injection data and Knudsen diffusion permeability model comparison.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

50 500 5000

k/k

0

Gas Pressure (psi)

Modeling

3.91mm data

5.44mm data

Page 178: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

163

Table 8 Apparent permeability measurement record of Marcellus shale thin disc samples using

pulse-decay method.

Execution

Sequence Sample

Gas Hydrostatic Pressure Repeated

Type Stress (psi) Range (psi) Cycles

1

5.39 mm

Marcellus

Shale

He 1600 64~1306 2

2

5.39 mm

Marcellus

Shale

He 3000 70~1970 2

3

5.39 mm

Marcellus

Shale

CO2 1600 55~1200 2

4

5.39 mm

Marcellus

Shale

CO2 3000 60~1484 2

5

3.91 mm

Marcellus

Shale

He 1600 65~1425 2

6

3.91 mm

Marcellus

Shale

He 3000 77~2025 2

7

3.91 mm

Marcellus

Shale

CO2 1600 64~1408 2

8

3.91 mm

Marcellus

Shale

CO2 3000 64~2014 2

Page 179: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

164

Chapter 6 Modeling of Apparent Permeability for Ultra-Tight Anthracite

and Shale Matrix: A Multi-mechanistic Flow Approach

Abstract

For a typical CBM/shale gas well, the late-time production rate appears to be steady

and indicates the sustainable long-term gas recovery potential of reservoirs (Lee et al.,

2014). Gas permeability is a critical and controlling parameter for the predictions of shale

gas and coalbed methane (CBM) productions. Shale matrix and tight anthracite are

characterized by ultra-tight pore structure and low permeability at micro- and nano-scale

with gas molecules stored by adsorption. Gas transport in shale and anthracite matrices no

longer always fall into the continuum flow regime described by Darcy’s law, rather a

considerable portion of transport is sporadic and irregular due to the mean free path of gas

becomes comparable to the prevailing scale. Therefore, this gas transport in anthracite coal

and shale will be a complicated nonlinear multi-mechanistic process influenced by non-

Darcy flow components like gas ad-/de-sorption, gas slippage and diffusion flow

(Javadpour, 2009; Azom and Javadpour, 2012; Zhang et al. 2012; Heller and Zoback,

2014). The complexity of the gas storage and flow mechanisms in ultra-fine pore structures

is diverse and makes it more difficult to predict the matrix permeability and gas

deliverability.

In this proposed study, a multi-mechanistic apparent-permeability model for

unconventional reservoir rocks (shale and coal) was extended from an existing Darcy

permeability model under proper stress condition (constant-stress and uniaxial-strain) (Liu

and Harpalani, 2013; Ma et al., 2011). This new model incorporates the pressure-dependent

Page 180: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

165

weighting coefficient to separate the contributions of Knudsen diffusion and Darcy flow

on shale and coal permeability. A combination of both permeability components was

coupled with pressure-dependent weighting coefficients. The final modeling result leads to

a good match with Pennsylvania anthracite and Marcellus shale sample permeability data.

The transformation of this model under hydrostatic condition was derived for the

convenience of future data analysis in different scenarios.

1. Introduction

Gas production from unconventional reservoirs such as gas shale and coalbed methane

(CBM) has become a major source of energy in the United States. The nature of tight

structures of coal/shale formations introduces various flow mechanisms which lead to a

more complex assessment of gas transport in tight reservoir rocks. The conceptual gas

transport process in shale reservoirs is a multi-mechanistic process, including both Darcy

and non-Darcy flow components. Several preceding analytical models have been proposed

to predict the evolution of apparent permeability for micro-porous media under various

boundary conditions (Harpalani and Chen, 1997; Javadpour, 2009; Civan et al., 2011; Liu

et al., 2010, 2012; Liu and Harpalani, 2013b; Ma et al., 2011; Wang et al., 2012). These

models demonstrate that the evolution of apparent permeability depends on the effective

stress and flow-induced micro-structuring (matrix deformation). On the other hand, some

other research efforts to date have involved the shale gas flow mechanism and shale gas

permeability based on granular nano-tube flow, and these models are summarized in Table

9. In Javadpour (2009), a capillary tube gas flow model was developed for shale to account

for Knudsen diffusion and slip flow using the slip boundary condition, which was

Page 181: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

166

combined into Darcy flow to describe the apparent permeability. Darabi et al. (2012)

proposed a new apparent permeability model considering Knudsen diffusion and surface

roughness to get a better flow transport prediction. Although these apparent permeability

models combined Darcy/continuum flow, slip flow and Knudsen diffusion components,

their flow contribution weighting factors are always constant, incapable of representing

correct multi-mechanistic gas transport phenomenon. In 2014, an apparent permeability

model was developed by Li et al. (2014), considering a higher flow regime coverage

including ad-/desorption, slip flow, transition flow, Knudsen diffusion and Darcy viscous

flow. Wu et al. (2014) presented a new apparent permeability model combining both Darcy

viscous flow and Knudsen diffusion with dynamic weighting factor. This factor is a

function of Knudsen number, capable of differentiating the contribution of different flow

regimes during gas pressure change. All the above permeability models provided good

fundamental understanding to predict the complex gas transport process and apparent

permeability. However, they start only from nano-scale molecular flow without proper

stress-strain boundary condition which made the model unrepresentative for gas transport

under actual reservoir condition. Therefore, transformative research and a better theoretical

model is still needed to link multi-mechanistic flow mechanisms to the apparent

permeability evolution and the overall gas deliverability under best-replicated reservoir

condition. This extended knowledge will provide an understanding of the complex fluid

dynamics in nano- and micro-pore inclusion porous media and its impact on the overall gas

deliverability in shales.

Page 182: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

167

2. Background and Theories

2.1 Diffusion and Molecular flow model

For completeness, we include the free molecular regime, which is relevant for gas-

phase transport at ambient conditions. The mass-flow rate of the free molecular regime was

modeled by Knudsen (1909). Roy et al. (2003) showed that Knudsen diffusion in nanopores

can be written in the form of pressure gradient. Gas mass flux by diffusion with negligible

viscous effects in a nanopore is described as:

𝐽𝐷 = 𝐷𝐾𝑀Δ𝑝

𝑅𝑇𝐿 (42)

where q is the flow rate, Δ𝑝 is the pressure drop across the rock sample, L is the length of

media and M is molar mass. Also, Knudsen diffusion constant 𝐷𝐾 is defined as:

𝐷𝐾 =2𝑟

3√

8𝑅𝑇

𝜋𝑀 (43)

In this Knudsen diffusion term, the porosity/tortuosity factor (𝜙/𝜏) is also introduced

to model Knudsen flow through porous media (Javadpour et al., 2007; Swami and Settari,

2012). The factor is to consider porous media as consisting of a certain percentage of open

pores (porosity) of the mean pore diameter and having a degree of interconnection resulting

in paths longer than a direct path (tortuosity) (Javadpour et al., 2007). In addition, the

fractal dimension of the pore surface is included to consider the effect of pore-surface

roughness on the Knudsen diffusion coefficient (Coppens, 1999; Coppens and Dammers,

2006). Surface roughness is one example of local heterogeneity. Increasing surface

roughness leads to an increase in residence time of molecules in porous media and a

decrease in Knudsen diffusivity. And this fractal dimension of the pore surface is a

Page 183: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

168

quantitative measure of surface roughness that varies between 2 and 3, representing a

smooth surface and a space-filling surface, respectively (Coppens and Dammers, 2006).

Then the modified Knudsen diffusion constant becomes:

𝐷𝐾′ = (𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏√

8𝑅𝑇

𝜋𝑀 (44)

where 𝐷𝑓 is the fractal dimension of the pore surface, and 𝛿′ is the ratio of normalized

molecular size 𝑑𝑚 to local average pore diameter 𝑑𝑝 , yielding 𝛿′ = 𝑑𝑚/𝑑𝑝 . Then we

substitute Equation (3) into (1):

𝐽𝐷 = 𝐷𝐾′ MΔ𝑝

𝑅𝑇𝐿= (𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏

MΔ𝑝

𝑅𝑇𝐿√

8𝑅𝑇

𝜋𝑀 (45)

2.2 Darcy/viscous flow model

For conventional reservoirs, since pore radius is in micrometers, Knudsen number is

very low (𝐾𝑛 < 0.01) (Karniadakis et al., 2005). The mean free path is negligible compared

to extremely small pore throat size. The flow is thus governed by viscous flow and driven

by pressure gradient. For this range of Knudsen number, the intermolecular collisions are

important and molecules’ collisions with pore walls are negligible (Swami et al., 2012). In

Choi et al. (2001), Hagen-Poiseuille equation was used to describe the molar flux of

viscous flow:

𝐽𝑉 = −𝜙

𝜏

𝑟2Δ𝑝

8𝜇𝑅𝑇𝐿 (46)

where 𝜇 is gas dynamic viscosity, which defines the diffusion of momentum due to the

intermolecular collisions. This dynamic viscosity should be corrected to match the varying

Page 184: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

169

rarefaction effect, and thus effective viscosity is applied and given as (Karniadakis et al.,

2005):

𝜇𝑒 = 𝜇(1

1+𝛼𝐾𝑛) (47)

where, α is the rarefaction coefficient.

3. Proposed Model

3.1 Apparent permeability model with multiple flows

At low pressure with small pore size, collisions are mainly between molecules and

walls. Knudsen diffusion flux depends on pressure gradient can be obtained by Equation

(6) (Choi et al., 2001; Roy et al., 2003; Coppens and Dammers, 2006). In order to model

the gas flow permeability and compare to laboratory data, such flux term needs to convert

into permeability term. Note that the relationship between mass flux and permeability in

continuous flow model can be determined as (Javadpour, 2009; Sakhaee-Pour and Bryant,

2012; Wu et al., 2014):

𝐽 =𝑘𝜌

𝜇∇𝑝 (48)

In this way we are able to obtain gas apparent permeability from the gas mass flux.

And the apparent permeability of Knudsen diffusion can be shown as:

𝐾𝐷 =𝐽𝐷𝜇

𝜌∇𝑝= (𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏

𝜇M

𝜌𝑅𝑇√

8𝑅𝑇

𝜋𝑀 (49)

For micro-scale flows such as transition flow and slip flow, a total mass flux of a gas

through a nanopore can be expressed as the result of a combination of weighted Darcy

viscous (Poiseuille) flow and Knudsen diffusion shown as (Javadpour, 2009):

𝐽𝑇 = 𝑊𝑉𝐽𝑉 + 𝑊𝐷𝐽𝐷 (50)

Page 185: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

170

where 𝑊𝑉 and 𝑊𝐷 are the weighting coefficient for viscous flow and Knudsen diffusion

respectively.

Similarly, a total gas apparent permeability can be also expressed as:

𝑘 = 𝑊𝑉𝐾𝑉 + 𝑊𝐷𝐾𝐷 (51)

The two weighting coefficients represent two complementary part of flow

contribution. They are modeled by following the basic mechanism that how the flow

regimes are classified, which is the frequency of gas molecule collisions with either nano-

pore walls or with each other. The contribution of viscous flow and Knudsen diffusion to

total gas transport can be determined by the ratio of molecules collision frequency and

nanopore wall-molecule collision frequency to overall collision frequency (Wu et al.,

2014). For example, by counting the ratio of the frequency that the gas molecules are

colliding with each other we can get at what percent the viscous flow is taking place.

Firstly, the number of collision appear within unit time is:

1

𝑡𝑡=

1

𝑡𝑚+

1

𝑡𝑠 (52)

where, 𝑡𝑡 is the average time consumed for one collision of overall gas molecules, 𝑡𝑚 is

the average time required for one collision between the molecules, and 𝑡𝑠 is the average

time required for one collision between nano-pore wall and molecule. And thus the number

of collision can also be expressed as (Reif.F, 1965):

1

𝑡𝑡=

𝑣𝑚

𝜆𝑇,

1

𝑡𝑚=

𝑣𝑚

𝜆,

1

𝑡𝑠=

𝑣𝑚

2𝑟 (53)

Page 186: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

171

where, 𝜆𝑇 is the average free path of the overall gas molecules, and 𝑣𝑀 is the average

velocity of molecule movement. By substitute Eqs. (12) into (11), we obtain the summation

of collisions in terms of different gas mean free path:

1

𝜆𝑇=

1

𝜆+

1

2𝑟 (54)

Note that the Knudsen number can be defined as:

𝐾𝑛 =𝜆

2𝑟 (55)

So the weighting coefficients for viscous flow and Knudsen diffusion can be derived

by combining Eqs. (12) ~ (14), shown as (Wu et al., 2014):

𝑊𝑉 =𝜆𝑇

𝜆=

1

(1+𝐾𝑛

2) (56)

𝑊𝐷 =𝜆𝑇

2𝑟=

1

(1+2

𝐾𝑛) (57)

By substituting Eqs. (4), (8), (15) and (16) in (10), the following apparent

permeability model can be obtained:

𝑘 = −1

(1+𝐾𝑛

2)

𝜙

𝜏

𝑟2

8𝑅𝑇𝜌+

1

(1+2

𝐾𝑛)

(𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏

𝜇M

𝜌𝑅𝑇√

8𝑅𝑇

𝜋𝑀 (58)

Since these two weighting coefficients are functions of Knudsen number, and thus

functions of pore pressure and pore diameter. The relative contribution of these two

coefficients are shown in Figure 57. The ratio becomes greater when pore pressure

decreases, indicating the significant contribution of Knudsen diffusion and slip flow effect

Page 187: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

172

at low pressure. Meanwhile, with smaller gas flow path, the more impact of Knudsen

weighting coefficient.

With the contribution of Darcy viscous flow and properly weighted by Knudsen

number dependent weighting factors, this gas apparent permeability model is able to

represent the gas flow among all the popular flow regimes within reasonable pressure and

pore size range, assuming this gas permeability only depends on flow-based parameters.

The comparison between the results of apparent permeability model and laboratory shale

sample permeability data collected in Chapter 5 for helium and CO2 are shown in Figure

58 and 59. Marcellus shale discs were used on step-wised pulse-decay test to carry out the

permeability data. Both Darcy flow and diffusion flow can be observed indirectly from

“check” shape permeability profiles by careful data analysis. Different from Knudsen

permeability (Wang and Liu, 2016), the permeability trend of this model are changing with

average pore size. Therefore, different pore sizes within 5-500 nm was specified to cover

the possible predicted curves. Compared to Knudsen diffusive permeability modeling

results, this apparent permeability model has a much better agreement with lab data. The

decline rate of permeability ratio is inversely proportional to the average pore size. For

helium modeling result in Figure 60, it matches the laboratory data best when the pore size

is about 25nm, and also works well for CO2 results. This gives a rough pore size baseline

of tested shale sample purely by modeling results, and indicates that the current modeling

result can be controlled by pore size dependent parameter. In Figure 60, the pore size

parameter in this flow model was adjusted to fit the apparent permeability value calculated

at each pressure step. Each fitted curve shows an agreement with experimental data for

Page 188: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

173

both sample, which gives evidence that this combined flow model can be used to predict

helium permeability evolution trend only if the pore size variable is properly input. Figure

61 shows the pore size distribution applied in screening process. We can clearly observe

that the overall trend of pore size for both sample is positive correlated to gas pressure. At

higher pressure, lower effective stress releases shale matrix and results in pores opening up

and thus larger average pore size. So we expect higher permeability values with increasing

pressure. With the contribution of viscous flow, this model makes better prediction than

single Knudsen apparent permeability model. It enhances the participation of Darcy flow

elements and thus able to also enhance the total permeability. However, for CO2

permeability, this model appears losing the track of increasing permeability trend

compared to laboratory data. According to Figure 61, the boundary permeability curves

represent average pore size at 5nm and 500nm both are positioning significantly below the

experimental data profile at high pressure. Even if a variable pore size parameter can be

applied, this model is still incapable of simulate CO2 permeability when reservoir pressure

is high and effective stress is relatively low. Although this model may be good enough to

cover different flow regimes by introducing proper flow parameters, it is believed still

lacking critical stress-strain boundary type controlling factors that better explain how the

permeability change with effective stress and matrix deformation at higher pressure range

(> 600 psi). If proper stress-strain condition is applied in this purely flow-based apparent

permeability model, we may expect a turning point and increasing trend at high pressure

representing the change of large-scale Darcy flow with effective stress and matrix

deformation.

Page 189: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

174

In order to develop a promising apparent permeability model, the contribution of

unconventional rock properties and mechanical effects need to be coupled into the existing

apparent permeability model. For Darcy permeability, the coal permeability model with

stress-strain condition that can better represent the Darcy flow permeability in situ reservoir

condition was introduced. As to non-Darcy flow, Knudsen diffusion was assumed to be the

main flow mechanism in unconventional gas transport when the gas pathway is extremely

narrow. Thus the Knudsen mass flux in nanopore based on molecular flow principles can

take the form of Darcy’s permeability and account for the contribution of Knudsen (non-

Darcy) permeability.

3.2 Apparent permeability model under in situ condition

To make good prediction of coal and shale reservoir production, various permeability

models representing reservoir in situ condition were presented and evolved. A proper

Darcy permeability model with stress-strain condition will lead to a reasonable

incorporation of Darcy viscous flow component into apparent permeability model.

Based on the relative orientation of coalbeds and cleats, Seidle et al. (1992)

represented the geometry of a coalbed reservoir as a collection of matchsticks, with each

matchstick representing a coal matrix block and void space between the sticks representing

the cleat volume. Using this matchstick model, Seidle et al. (1992) then proposed a model

to predict permeability of coal in terms of changes in stress resulting from continued

production as:

𝑘

𝑘0= exp [−3𝐶𝑓(𝜎 − 𝜎0)] (59)

Page 190: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

175

where, 𝑘0 is initial permeability, 𝐶𝑓 is cleat compressibility, 𝜎 is the hydrostatic stress and

𝜎0 is initial hydrostatic stress. However, this model was formed without addressing proper

stress-strain type boundary condition.

Shi and Durucan, (2005) proposed a uniaxial-strain permeability model (S&D

model), also based on matchstick model, where cleat permeability is impacted by the

prevailing effective horizontal stresses acting normal to the cleats. Following the effective

stress-strain relationship the variation in effective horizontal stress under uniaxial strain

condition was expressed as a function of pore pressure as follows:

𝜎 − 𝜎0 = −𝑣

1−𝑣(𝑝 − 𝑝0) +

𝐸

3(1−𝑣)휀𝑙(

𝑝

𝑝+𝑃𝜀−

𝑝0

𝑝0+𝑃𝜀) (60)

where 𝑝0 is the initial pressure, E is young’s modulus, 휀𝑙 represents the maximum strain

which can be achieved at infinite pressure, and 𝑃𝜀 is the pressure at which coal attains half

of the maximum strain. Taking the form of effective stress, this model is easy to be

incorporated into the permeability model by Seidle et al. (1992). However, this model

showed a poor agreement when pressure is lower than 2.1 MPa and was lacking scientific

improvement.

A modification of S&D model was then proposed by (Liu et al., 2012). Following

stress-strain relationship for isothermal gas desorbing process on coal, the modified model

corrected S&D model made a direct analogy between thermal contraction and matrix

shrinkage under uniaxial-strain condition, shown as:

𝜎 − 𝜎0 = −𝑣

1−𝑣(𝑝 − 𝑝0) +

𝐸(1+𝑣)

3(1−𝑣)휀𝑙(

𝑝

𝑝+𝑃𝜀−

𝑝0

𝑝0+𝑃𝜀) (61)

Page 191: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

176

Combining Equation (18) and (20), we can obtain the modified S&D model in terms

of Darcy permeability:

𝑘

𝑘0= exp [−3𝐶𝑓 (−

𝑣

1−𝑣(𝑝 − 𝑝0) +

𝐸(1+𝑣)

3(1−𝑣)휀𝑙(

𝑝

𝑝+𝑃𝜀−

𝑝0

𝑝0+𝑃𝜀))] (62)

3.3 Sorption-induced stress-strain model

Some empirical approaches have been developed to establish a link between the

measured strain and reservoir pressure to quantitatively model the sorption-induced

volumetric strain. Levine (1996) used a Langmuir-type model to fit the experimentally-

derived linear strain data for Illinois basin coals, when flooded with methane and CO2:

휀 = 휀𝑔𝑉𝐿𝑝

𝑝+𝑃𝐿 (63)

where 휀 is the sorption-induced volumetric strain at pressure p, and 휀𝑔 represents the

coefficient of sorption-induced volumetric strain.

With the help of this empirical model, a number of analytical permeability models

were developed to quantitatively incorporate the shrinkage effect (Izadi et al., 2011; Ma et

al., 2011; Wang et al., 2012a). Liu and Harpalani (2013a) proposed a theoretical technique

to model the volumetric changes in coal matrix during gas de/ad- sorption using elastic

properties, sorption parameters and physical properties of coal. The proposed model is

based on the principles of physics and chemistry of a surface and the interface theory. The

volumetric strain of coal matrix for sorbing gas includes two components, sorption-induced

and mechanical-induced strain. The sorption-induced strain is directly proportional to the

decrease in surface energy and mechanical-induced strain is calculated by the Hooke's law.

Page 192: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

177

In this model, these two strains were assumed to be purely additive. It highlights the

relationships between sorption and coal matrix strain. The volumetric strain is calculated

as follows:

휀 =3𝑉𝐿𝜌𝑠𝑅𝑇

𝐸𝐴𝑉0∫

1

𝑝+𝑃𝐿𝑑𝑝

𝑝

0−

3(1−2𝜐)

𝐸∫ 𝑑𝑝

𝑝

0 (64)

where 𝜌𝑠 is the density of the solid adsorbent (coal), 𝐸𝐴 is the modulus of the solid

expansion, E is the Young's modulus, 𝜐 is the Poisson's ratio, 𝑉0 is the gas molar volume

(22.4 m3/kmol).

In order to couple this model with different permeability models, the stress-

dependent-strain component must be dropped for two reasons. First, the mechanical

compression of coal is not analogous to thermal shrinkage since the two mechanisms are

entirely different. Second, the permeability models already take this component into

account and, by including this, it would effectively be counted twice. Therefore, the

sorption-induced strain alone for different pressures can be given as:

Δ휀 =3𝑉𝐿𝜌𝑠𝑅𝑇

𝐸𝐴𝑉0∫

1

𝑝+𝑃𝐿𝑑𝑝

𝑝2

𝑝1 (65)

where Δ휀 is the sorption-induced incremental volumetric strain, p1 and p2 are reservoir

pressures at conditions one and two.

Equation (21) can be thus rewritten as the following equation to incorporate the

sorption-induced strain model in Equation (24):

𝑘

𝑘0= exp [−3𝐶𝑓 (−

𝑣

1−𝑣(𝑝 − 𝑝0) +

𝐸(1+𝑣)

3(1−𝑣)Δ휀)] (66)

Page 193: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

178

Therefore, the apparent permeability for nanopore gas flow transformed from

Equation (10) becomes:

𝑘 = 𝑊𝑉𝑘 + 𝑊𝐷𝐽𝐷𝜇

𝜌∇𝑝 (67)

Rearrange Equation (26) in terms of permeability ratios, we can get:

𝑘

𝑘0= 𝑊𝑉

𝑘

𝑘0+ 𝑊𝐷

𝐽𝐷𝜇

𝜌∇𝑝𝑘𝐷0 (68)

By substitute Equation (21), (26), (30), (36) and (37) into (41), we can obtain the

intact form of proposed apparent permeability model:

𝑘

𝑘0=

1

(1+𝐾𝑛

2)

exp [−3𝐶𝑓 (−𝑣

1−𝑣(𝑝 − 𝑝0) +

𝐸(1+𝑣)

3(1−𝑣)Δ휀)] +

1

(1+2

𝐾𝑛) (𝛿′)𝐷𝑓−2 2𝑟𝜙

3𝜏𝑘𝐷0

μM

𝜌𝑅𝑇√

8𝑅𝑇

𝜋𝑀 (69)

3.4 Permeability model under hydrostatic stress condition

The above apparent permeability model is developed for permeability under uniaxial

strain condition only. However, in numerous laboratory cases, permeability data were

collected by applying the rock sample hydrostatic condition or constant stress condition.

The proposed model is expected to have the capacity to predict permeabilities not only

under one scenario but also for the convenience of analyzing various cases. Therefore, in

this section, stress–strain relationships for a linear elastic gas-desorbing porous medium

are derived. The derivation makes use of an analogy between thermal contraction and

matrix shrinkage associated with gas desorption in coalbeds. Stress– strain relationships

for a thermoelastic porous medium can be readily found in the literature. Replacing the

Page 194: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

179

thermal expansion term by an analogous shrinkage term yields the corresponding

relationships for an isothermal gas-desorbing porous medium. We start the derivation with

the (effective) stress–strain relationship for a homogenous, isotropic thermoelastic porous

medium (Shi and Durucan, 2004; Liu et al., 2012) :

�̃�𝑖𝑗 = 2𝐺휀�̃�𝑗 + 𝜆휀̃𝛿𝑖𝑗 + (𝜆 +2

3𝐺)𝛼𝑇�̃�𝛿𝑖𝑗 (70)

휀̃ = 휀�̃�𝑥 + 휀�̃�𝑦 + 휀�̃�𝑧 (71)

�̃�𝑖𝑗 = �̃�𝑖𝑗 + 𝑝𝛿𝑖𝑗 (72)

where the sign ∼ indicates incremental values, G and λ are the Lame constants of the porous

rock, 𝛿𝑖𝑗 is the Kronecker delta, ε is volumetric strain, 휀𝑖𝑗 is linear strain, T is temperature,

𝜎 is total stress, 𝜏 is total stress, p is pore pressure and 𝛼𝑇 is the coefficient of volumetric

thermal expansion.

In a non-isothermal body, if the temperature drops the fabric shrinks, leading to an

increase in the porous medium porosity. This is directly analogous to matrix shrinkage in

coalbeds, where cleat porosity increases as gas desorbs during draw-down (Palmer and

Mansoori, 1998). By making a direct analogy between thermal contraction and matrix

shrinkage, stress–strain relationships for an isothermal gas-desorbing rock may be written

as (positive in compression):

�̃�𝑖𝑗 = 2𝐺휀�̃�𝑗 + 𝜆휀̃𝛿𝑖𝑗 + (𝜆 +2

3𝐺)휀�̃�𝛿𝑖𝑗 (73)

where 휀𝑠 is the macroscopic volumetric matrix shrinkage strain induced by gas desorption

from coal. The stress–strain relationships for the three normal stress components are of

Page 195: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

180

interest here, as the shear stress components are not affected by thermal contraction/matrix

shrinkage.

�̃�𝑥𝑥 = 2𝐺휀�̃�𝑥 + 𝜆휀̃ + (𝜆 +2

3𝐺)휀�̃� (74)

�̃�𝑦𝑦 = 2𝐺휀�̃�𝑦 + 𝜆휀̃ + (𝜆 +2

3𝐺)휀�̃� (75)

�̃�𝑧𝑧 = 2𝐺휀�̃�𝑧 + 𝜆휀̃ + (𝜆 +2

3𝐺)휀�̃� (76)

For hydrostatic conditions, the stresses and strains of all direction should follow �̃�𝑥𝑥 =

�̃�𝑦𝑦 = �̃�𝑧𝑧 and 휀�̃�𝑥 = 휀�̃�𝑦 = 휀�̃�𝑧 . Then the following relationship was formulated,

assuming the changes of stress/strain in 3 directions to be the same:

�̃�𝑥𝑥 = �̃�𝑦𝑦 = �̃�𝑧𝑧 = (2𝐺 + 3𝜆)휀�̃�𝑥 + (𝜆 +2

3𝐺)휀�̃� (77)

Assuming coal is a linearly elastic and isotropic material. The deformation of coal

(strains) can then be calculated for known increments of stress if two coal elastic constants

are specified:

휀�̃�𝑥 =�̃�𝑥𝑥

𝐸−

𝜈

𝐸�̃�𝑦𝑦 −

𝜈

𝐸�̃�𝑧𝑧 =

1−2𝜈

𝐸�̃�𝑥𝑥 (78)

Combining Equation (36) and (37),

�̃�𝑥𝑥 = �̃�𝑦𝑦 = �̃�𝑧𝑧 = (2𝐺 + 3𝜆)1−2𝜈

𝐸�̃�𝑥𝑥 + (𝜆 +

2

3𝐺)휀�̃� (79)

Since the applied stresses on 3 directions are constant during pressure change under

hydrostatic stress condition, the total stress increment on x, y, z axis always remains 0:

�̃�𝑥𝑥 + 𝑝 = �̃�𝑦𝑦 + 𝑝 = �̃�𝑧𝑧 + 𝑝 = 0 (80)

Now Equation (38) becomes:

�̃�𝑥𝑥 = �̃�𝑦𝑦 = �̃�𝑧𝑧 = −(2𝐺 + 3𝜆)1−2𝜈

𝐸𝑝 + (𝜆 +

2

3𝐺)휀�̃� (81)

Page 196: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

181

Recalling that the relationship between E, G, ν and 𝜆 for homogeneous and isotropic

coal:

𝐺 =𝐸

2(1+𝜈) (82)

𝜆 =𝐸𝜈

(1+𝜈)(1−2𝜈) (83)

By substitute Equation (41) and (42) in (40):

�̃�𝑥𝑥 = �̃�𝑦𝑦 = �̃�𝑧𝑧 = −𝑝 +𝐸

3(1−2𝜈)휀�̃� (84)

If we combine Equation (18) and (43), the final coal permeability can be written as:

𝑘

𝑘0= exp [−3𝐶𝑓 (−(𝑝 − 𝑝0) +

𝐸

3(1−2𝑣)Δ휀)] (85)

Therefore, under hydrostatic condition, Equation (28) will become:

𝑘𝑎

𝑘𝑎0=

1

(1 +𝐾𝑛

2 )exp [−3𝐶𝑓 (−

𝑣

1 − 𝑣(𝑝 − 𝑝0) +

𝐸

3(1 − 2𝑣)Δ휀)]

+1

(1 +2

𝐾𝑛)

(𝛿′)𝐷𝑓−22𝑟𝜙

3𝜏𝑘𝐷0

μM

𝜌𝑅𝑇√

8𝑅𝑇

𝜋𝑀 (86)

4. Data validation and discussion of proposed apparent permeability model

In this work, the propose apparent permeability model, combining modified S&D

permeability model (Darcy component) and Knudsen permeability (non-Darcy

component), was validated by the laboratory coal permeability data collected from Wang

et al. (2015) and shale permeability data from Wang and Liu (2017). All the required input

parameters were estimated based on the reported Pennsylvania anthracite and Marcellus

shale rock sample properties (Wang et al., 2015; Wang and Liu, 2016). The modeled results

Page 197: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

182

and experimental data of both anthracite and shale sample permeability were concluded

from Figure 62 to 65. First of all, for the anthracite permeability with helium injection, the

proposed model coupled with Liu et al.’s model has a relatively poor agreement comparing

to that under hydrostatic condition shown in Figure 62. Although the permeability data

were not measured under exact hydrostatic condition, the unchanged stress and strain

incremental values during pressure variation are also zero for constant stress scenario. In

this case, the proposed hydrostatic model is still principally valid and brought an almost

perfect match with anthracite permeability data. The discrepancy between the matching

results of two models originates from how these two stress-strain conditions primarily

influence sample permeability during pulse-decay test. During pressure depletion, the

increase of effective horizontal stress for uniaxial strain condition is slower than for

constant stress condition. The reason is simply because horizontal stress under uniaxial

strain condition kept decreasing to maintain the zero horizontal strain resulting in a

relatively slow effective horizontal stress increase (Wang et al., 2015). The permeability

under uniaxial strain condition suffers from less intensive compress from effective stress

and usually tends to be larger than in other cases. Consequently, the modeling result of

coupled Liu et al.’s model are expected to be larger than the actual data measured in

constant stress condition. This is a consistent phenomenon that can be observed from all

the modeling results. On the other hand, the availability of hydrostatic model and the failure

of uniaxial strain model for validating helium anthracite permeability data under constant

stress condition prove the validity of our proposed model.

Page 198: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

183

As to sorbing gas, we have been expecting “check” shape permeability profile from

either laboratory data or modeling results (Wang et al., 2015). In Figure 63, the proposed

apparent permeability model shows its capability of replicating such behavior by

addressing sorption-induced strain effect in Darcy permeability component. The sorption-

induced strain term has positive value and, together with the significant Knudsen diffusion

effect, enhanced the low pressure CO2 permeability. Both models are showing similar

“check” shape trends except the discrepancy between two conditions we discussed above.

The hydrostatic model still presents a much better match with CO2 experimental data than

uniaxial strain model does, with an average deviation of 5%~8%.

The primary target reservoir rocks of this proposed model are unconventional source

rock, such as coal and shale. This proposed model can be considered a combination of coal

permeability model and shale gas flow model, so its viability on both coal and shale needs

to be validated. Based on the results from Figure 62 and 63, we can conclude that this

proposed model is suitable for coal permeability prediction. However, we still have doubt

on its availability on shale permeability since the primary Darcy component, which

includes entire mechanical term, derives from coal models only. Therefore, we here

introduce the assumption that this proposed model may also provide good perdition of shale

permeability data since shale is also extremely tight in structure as some types of coal. And

a group of parallel validation was also conducted by using shale disc permeability data and

shown in Figure 64 and 65. Overall, both uniaxial strain model and hydrostatic model have

relatively good fit with little deviation from original data. Theoretically, there would be a

difficulty for the proposed model to predict these shale data. That is: the shale permeability

Page 199: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

184

data are collected by conducting pulse-decay test on very thin shale discs (around 5mm in

thickness). Although the stress condition was manually maintained hydrostatic, it is hard

to prove whether effective stress still has considerable influence on permeability evolution

or not. From Figure 64, we can see at low pressure (<2MPa) there is a helium permeability

turning up and thus create a “check” shape just as CO2 data did. This results from the

extremely tight shale structure and the significantly enhanced Knudsen diffusion effect

(non-Darcy), since Knudsen permeability can be much higher at low pressure inside tiny

pores (Wang and Liu, 2016). Fortunately, the proposed model under both conditions can

successfully reflect such behavior and still has good match with experimental data. For

CO2 data in Figure 65, the matching results are just similar to helium results. These

repeated validation works under different cases have provided good proof on testing the

capacity of our proposed apparent permeability model. In future work, we may need long

shale core sample for permeability measurements that can eliminate the uncertainty in

applying stresses. Thus, the implement of this proposed model on shale permeability

evolution can be further justified.

5. Conclusions

1. Since permeability can vary significantly with effective stress at high pressure, stress-

strain type boundary conditions is critical in modeling unconventional rock

permeability evolution. Otherwise, the turning point when the Darcy permeability starts

to dominate and increase with decreasing effective stress is difficult to represent.

2. By varying pore size, which results from the change of effective stress and sorption-

Page 200: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

185

induced strain, we may obtain the correct fit of permeability data from nanotube flow-

based model. This explains that knowing rock structure change is critical for determine

the promising apparent permeability model with multiple flow mechanisms.

3. The proposed apparent permeability model under both uniaxial strain condition and

hydrostatic condition potentially have good capacity on unconventional reservoir rock

permeability prediction. The hydrostatic model showed good fit with anthracite

permeability data collected under constant stress condition.

4. The prediction of shale permeability by this proposed model is promising based on

good data validation results. But further efforts are needed on solid shale core

permeability data measurements under correctly stressed condition.

References

Choi, J.-G., D. D. Do, and H. D. Do, 2001, Surface Diffusion of Adsorbed Molecules in

Porous Media: Monolayer, Multilayer, and Capillary Condensation Regimes:

Industrial & Engineering Chemistry Research, v. 40, no. 19, p. 4005–4031,

doi:10.1021/ie010195z.

Civan, F., C. Rai, and C. Sondergeld, 2011, Shale-gas permeability and diffusivity

inferred by improved formulation of relevant retention and transport mechanisms:

Transport in Porous Media, v. 86, no. 3, p. 925–944, doi:10.1007/s11242-010-9665-

x.

Coppens, M., 1999, The effect of fractal surface roughness on diffusion and reaction in

porous catalysts – from fundamentals to practical applications: Catalysis Today, v.

53, no. 2, p. 225–243, doi:10.1016/S0920-5861(99)00118-2.

Coppens, M. O., and A. J. Dammers, 2006, Effects of heterogeneity on diffusion in

nanopores - From inorganic materials to protein crystals and ion channels: Fluid

Phase Equilibria, v. 241, no. 1–2, p. 308–316, doi:10.1016/j.fluid.2005.12.039.

Darabi, H., a. Ettehad, F. Javadpour, and K. Sepehrnoori, 2012, Gas flow in ultra-tight

shale strata: Journal of Fluid Mechanics, v. 710, p. 641–658,

doi:10.1017/jfm.2012.424.

Page 201: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

186

Ertekin, T., G. King, and F. Schwerer, 1986, Dynamic Gas Slippage: A Unique Dual-

Mechanism Approach to the Flow of Gas in Tight Formations: p. 43–52,

doi:10.2118/12045-PA.

Harpalani, S., and G. Chen, 1997, Influence of gas production induced volumetric strain

on permeability of coal: Geotechnical and Geological Engineering, v. 15, no. 4, p.

303–325, doi:10.1007/BF00880711.

Izadi, G., and D. Elsworth, 2013, The effects of thermal stress and fluid pressure on

induced seismicity during stimulation to production within fractured reservoirs:

Terra Nova, v. 25, no. 5, p. 374–380, doi:10.1111/ter.12046.

Izadi, G., S. Wang, D. Elsworth, J. Liu, Y. Wu, and D. Pone, 2011, Permeability

evolution of fluid-infiltrated coal containing discrete fractures: International Journal

of Coal Geology, v. 85, no. 2, p. 202–211, doi:10.1016/j.coal.2010.10.006.

Javadpour, F., 2009, Nanopores and apparent permeability of gas flow in mudrocks

(shales and siltstone): Journal of Canadian Petroleum Technology, v. 48, p. 1–6.

Javadpour, F., D. Fisher, and M. Unsworth, 2007, Nanoscale gas flow in shale gas

sediments: Journal of Canadian Petroleum Technology, v. 46, no. 10,

doi:10.2118/07-10-06.

Karniadakis, G., A. Beskok, and A. Narayan, 2005, Microflows and Nanoflows: 817 p.,

doi:10.1007/0-387-28676-4.

Levine, J., 1996, Model study of the influence of matrix shrinkage on absolute

permeability of coal bed reservoirs: Geological Society, London, Special

Publications.

Li, Y., X. Li, J. Shi, H. Wang, L. Wu, S. Teng, and S. Corp, 2014, A Nano-Pore Scale

Gas Flow Model for Shale Gas Reservoir Flow characteristic of shale gas reservoir

in various pole scales and molecule scales: no. June, p. 9–11.

Li, J., D. Liu, S. Lu, Y. Yao, and H. Xue, 2015, Evaluation and modeling of the CO2

permeability variation by coupling effective pore size evolution in anthracite coal:

Energy & Fuels, doi:10.1021/ef502674y.

Liu, J., Z. Chen, D. Elsworth, X. Miao, and X. Mao, 2010, Linking gas-sorption induced

changes in coal permeability to directional strains through a modulus reduction

ratio: International Journal of Coal Geology, v. 83, no. 1, p. 21–30,

doi:10.1016/j.coal.2010.04.006.

Liu, S., and S. Harpalani, 2013a, A new theoretical approach to model sorption-induced

coal shrinkage or swelling: AAPG bulletin, v. 97, no. 7, p. 1033–1049,

doi:10.1306/12181212061.

Liu, S., and S. Harpalani, 2013b, Permeability prediction of coalbed methane reservoirs

during primary depletion: International Journal of Coal Geology, v. 113, p. 1–10,

doi:10.1016/j.coal.2013.03.010.

Page 202: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

187

Liu, S., S. Harpalani, and M. Pillalamarry, 2012, Laboratory measurement and modeling

of coal permeability with continued methane production: Part 2 – Modeling results:

Fuel, v. 94, p. 117–124, doi:10.1016/j.fuel.2011.10.053.

Ma, Q., S. Harpalani, and S. Liu, 2011, A simplified permeability model for coalbed

methane reservoirs based on matchstick strain and constant volume theory:

International Journal of Coal Geology, v. 85, no. 1, p. 43–48,

doi:10.1016/j.coal.2010.09.007.

Palmer, S. I., and J. Mansoori, 1998, How permeability depends on stress and pore

pressure in coalbeds: a new model: Soc Petrol Eng Res Eng, no. December, p. 539–

544.

Reif.F, 1965, Fundamentals of statistical and thermal physics.

Roy, S., R. Raju, H. F. Chuang, B. a. Cruden, and M. Meyyappan, 2003, Modeling gas

flow through microchannels and nanopores: Journal of Applied Physics, v. 93, no. 8,

p. 4870–4879, doi:10.1063/1.1559936.

Sakhaee-Pour, A., and S. Bryant, 2012, Gas Permeability of Shale: SPE Reservoir

Evaluation & Engineering, v. 15, p. 401–409.

Seidle, J. P., M. W. Jeansonne, and D. J. Erickson, 1992, Application of matchstick

geometry to stress dependent permeability in coals: SPE Rocky Mountain Regional

Meeting, p. 433–444, doi:10.2523/24361-MS.

Shabro, V., 2012, Forecasting Gas Production in Organic Shale with the Combined

Numerical Simulation of Gas Diffusion in Kerogen, Langmuir Desorption from

Kerogen Surfaces, and Advection in Nanopores: SPE Annual Technical …,

doi:10.2118/159250-MS.

Shi, J., and S. Durucan, 2005, A model for changes in coalbed permeability during

primary and enhanced methane recovery: SPE Reservoir Evaluation & Engineering,

v. 8, no. 4, p. 291–299.

Shi, J. Q., and S. Durucan, 2004, Drawdown Induced Changes in Permeability of

Coalbeds: A New Interpretation of the Reservoir Response to Primary Recovery:

Transport in Porous Media, v. 56, no. 1, p. 1–16,

doi:10.1023/B:TIPM.0000018398.19928.5a.

Singh, H., and F. Javadpour, 2015, Langmuir slip-Langmuir sorption permeability model

of shale: FUEL, no. October, doi:10.1016/j.fuel.2015.09.073.

Singh, H., and F. Javadpour, 2013, Nonempirical Apparent Permeability of Shale, in

Unconventional Resources Technology Conference: p. 16.

SUGGESTED METHODS FOR UNIAXIAL‐STRAIN COMPRESSIBILITY

TESTING FOR RESERVOIR GEOMECHANICS, 2015: v. c, no. April, p. 1–29.

Swami, V., C. Clarkson, and A. Settari, 2012, Non-Darcy Flow in Shale Nanopores: Do

We Have a Final Answer? SPE Canadian Unconventional …, no. November, p. 1–

17.

Page 203: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

188

Swami, V., and A. Settari, 2012, A pore scale gas flow model for shale gas reservoir, in

SPE Americas Unconventional Resources Conference held in Pittsburgh: p. 5–7.

Wang, S., D. Elsworth, and J. Liu, 2012, A mechanistic model for permeability evolution

in fractured sorbing media: Journal of Geophysical Research, v. 117, no. B6.

Wang, S., D. Elsworth, and J. Liu, 2012, A mechanistic model for permeability evolution

in fractured sorbing media: Journal of Geophysical Research: Solid Earth ( …, v.

117, no. B6, p. B06205, doi:10.1029/2011JB008855.

Wang, Y., and S. Liu, 2016, Estimation of Pressure-Dependent Diffusive Permeability of

Coal Using Methane Diffusion Coefficient: Laboratory Measurements and

Modeling: Energy and Fuels, v. 30, no. 11, p. 8968–8976,

doi:10.1021/acs.energyfuels.6b01480.

Wang, Y., S. Liu, and D. Elsworth, 2015, Laboratory investigations of gas flow behaviors

in tight anthracite and evaluation of different pulse-decay methods on permeability

estimation: International Journal of Coal Geology, v. 149, p. 118–128,

doi:10.1016/j.coal.2015.07.009.

Wu, K., X. Li, C. Wang, Z. Chen, and W. Yu, 2015, A Model for Gas Transport in

Microfractures of Shale and Tight Gas Reservoirs: AIChE Journal, v. 61, no. 6,

doi:10.1002/aic.

Wu, K., X. Li, C. Wang, W. Yu, and Z. Chen, 2014, Apparent Permeability for Gas Flow

in Shale Reservoirs Coupling Effects of Gas Diffusion and Desorption: Proceedings

of the 2nd Unconventional Resources Technology Conference, p. 1–18,

doi:10.15530/urtec-2014-1921039.

Page 204: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

189

Appendix

Figure 57 Contribution of Knudsen diffusion flow relative to Darcy viscous flow in terms of

weighting coefficient, as a function of pore pressure and pore radius.

0.0001

0.001

0.01

0.1

1

10

0 200 400 600 800 1000 1200 1400 1600

log

(W

D/W

V)

Pore Pressure (psi)

r=5 nm

r=25 nm

r=500 nm

Page 205: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

190

Figure 58 Shale He data and combined flow permeability model comparison.

0

0.2

0.4

0.6

0.8

1

1.2

0 500 1000 1500 2000

k/k

0

Pressure (psi)

rp=5nm

rp=25nm

rp=500nm

3.91mm disc data

5.44mm disc data

Page 206: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

191

Figure 59 Shale CO2 data and combined flow permeability model comparison.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 500 1000 1500 2000

k/k

0

Pressure (psi)

rp=5nm

rp=25nm

rp=500nm

3.91mm disc data

5.44mm disc data

Page 207: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

192

Figure 60 Combined flow model matching experimental data by varying pore size parameter in

apparent permeability model.

0

0.2

0.4

0.6

0.8

1

1.2

0 200 400 600 800 1000 1200 1400 1600

k/k

0

Pressure (psi)

3.91mm disc data

5.44mm disc data

rp=5nm

rp=500nm

Screen 3.91mm

Screen 5.44mm

Page 208: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

193

Figure 61 Pore size distribution applied in combined apparent permeability model for fitting

experimental data (screening process).

0

20

40

60

80

100

120

0 500 1000 1500 2000

Po

re S

ize

(nm

)

Pressure (psi)

3.91mm disc

5.44mm disc

Page 209: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

194

Figure 62 The proposed apparent permeability modeling results (He) with lab data validation for

natural-fractured anthracite core sample under uniaxial strain condition and hydrostatic condition.

0

0.2

0.4

0.6

0.8

1

1.2

0 200 400 600 800 1000

k/k

0

Pore Pressure (psi)

Experimental Data

Proposed Model Under Hydrostatic Condition

Proposed Model Coupled with Liu et al. 2012 Model

Page 210: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

195

Figure 63 The proposed apparent permeability modeling results (CO2) with lab data validation for

natural-fractured anthracite core sample under uniaxial strain condition and hydrostatic condition.

0

0.2

0.4

0.6

0.8

1

1.2

0 200 400 600 800 1000

k/k

0

Pore Pressure (psi)

Experimental Data

Proposed Model Under Hydrostatic Condition

Proposed Model Coupled with Liu et al. 2012 Model

Page 211: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

196

Figure 64 The proposed apparent permeability modeling results (He) with lab data validation for

shale disc samples under uniaxial strain condition and hydrostatic condition.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 200 400 600 800 1000 1200 1400

k/k

0

Pore Pressure (psi)

Experimental Data

Proposed Model Under Hydrostatic Condition

Proposed Model Coupled with Liu et al. 2012 Model

Page 212: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

197

Figure 65 The proposed apparent permeability modeling results (CO2) with lab data validation for

shale disc samples under uniaxial strain condition and hydrostatic condition.

0

0.2

0.4

0.6

0.8

1

1.2

0 200 400 600 800 1000 1200 1400 1600

k/k

0

Pore Pressure (psi)

Experimental Data

Proposed Model Under Hydrostatic Condition

Proposed Model Coupled with Liu et al. 2012 Model

Page 213: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

198

Table 9 Comparison of gas apparent permeability models

Model Description Comments

Ertekin et al. (1986)

Combination of continuum flow

and Fick diffusion with constant

contribution weight

Constant flow weighting factors

Javadpour (2009)

Combination of slip flow and

Knudsen diffusion with linear

additivity

Flow weighting factor is 1

Darabi et al. (2012)

Update from Javadpour (2009),

accounting for surface roughness

and Knudsen diffusion in a porous

medium

Ignored desorption process

Sakhaee-Pour and Bryant (2012)

Model developed using slip flow

assumption, represented by

Maxwell theory. Accounts for

Knudsen diffusion

Complex numerical model

Shabro (2012)

A finite-difference based

numerical model and geometrical

parameters are used to reconstruct

porous structure of shale

Complex numerical model only for ideal

gas

Singh and Javadpour (2013)

Model developed using Navier–

Stokes equation and kinetic theory

(no slip flow assumption).

Accounts for Knudsen diffusion,

porous medium and sorption

Ignored slip flow

Li et al. (2014)

Combination of continuum flow

and Knudsen diffusion with

contribution weight

has overlaps in flow regimes

Page 214: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

199

Wu et al. (2014)

Dynamic weighting coefficients to

separate the contribution of

Knudsen diffusion and Darcy

viscous flow

Lack of stress-strain type boundary

condition

Li et al. (2015)

Add the effective stress and matrix

shrinkage based on Javadpour’s

model, coupling the stress model

in pore radius parameter

Flow weighting factor is 1

Singh and Javadpour (2015)

Add sorption effect in Javadpour’s

model

Lack of stress-strain type boundary

condition

Wu et al. (2015)

Update of Wu et al. (2014) with

introducing fracture geometer

analysis

based on simplified microfracture model

Page 215: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

200

Chapter 7 Conclusions

The main purpose of this research is to develop a new pressure-dependent apparent

permeability model under different stress conditions, which can properly differentiate the

contribution of both Darcy and non-Darcy flow and thus can account for multi-mechanistic

gas transport process in unconventional reservoirs. A series of experimental studies and

theoretical analyses on adsorption isotherm, diffusion coefficient and gas permeability of

Hazelton anthracite core sample and Marcellus shale disc sample under different stressed

controlled conditions has been presented. The results show that the permeability evolution

of studied coal and shale is pressure-/stress- dependent and it is simultaneously controlled

by the effective stress profile and the sorption process which tends to alter the

microstructure of coal and shale. Three different pulse-decay calculation methods for

conventional and unconventional gas permeability are utilized and compared. Laboratory

measured diffusion coefficient was used to obtain an equivalent permeability so that the

Knudsen modeled results got independently compared and evaluated. The final modeling

result made a good agreement with experimental data. Based on the work completed, the

following conclusions are made and summarized:

1. Under uniaxial strain condition, the applied horizontal stress linearly decreased

with gas pressure depletion, which consists with traditional oil/gas reservoirs.

Because of the horizontal stress loss, the permeability under constant stress

condition with helium depletion was found to be less than the results under uniaxial

strain condition.

Page 216: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

201

2. For the pulse-decay method, the contribution of ad-/desorption can be clearly

observed from the pressure respond curves. And this effect is stronger at low

pressure than high pressures. The sorption induced matrix shrinkage plays

important role on the permeability enhancement at low pressure for the anthracite

coal. But it is not strong enough to compensate the stress effect as the bituminous

coal did. Cui et al.’s method gives an obvious enhancement for the contribution of

sorption effect in permeability calculation, providing a good direction of how the

gas sorption effect can help to predict apparent permeability.

3. The significance of diffusion flow depends on the ad-/desorption rate. At low

pressure (< 2MPa/280 psi) when the sorption process is severer, higher diffusion

coefficient values can be observed. Generally, a negative correlation between

diffusion coefficient and pressure is expected. The equivalent permeability

converted from diffusion coefficient can reflect the contribution of diffusion

process in CBM gas transport if assumed diffusion process dominates only. The

overall value of diffusive permeability is quite low. At low pressure (< 2MPa/280

psi), the diffusive permeability can match with preceding experience on coal gas

permeability since they all have increasing trends when pressure keep decreasing.

4. The sorption-induced linear strains are anisotropic for both methane and CO2.

Generally, the strain measured perpendicular to the bedding (┴) was greater than

parallel (//) to it. The CO2 sorption-induced strains were not fully reversible, but

they are for methane. The volumetric CO2 sorption induced strain was more than

three times that induced by methane. The application of isotropic deformation in

Page 217: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

202

permeability prediction model can overestimate the permeability loss compared to

anisotropic deformation. This demonstrates that the anisotropic coal deformation

should be further considered to predict the permeability behavior in the coal-CO2

sequestration project.

5. The main flow regimes are dynamically changing during shale gas production, and

all flow regimes including Darcy viscous flow, slip flow, transition flow and

Knudsen diffusion exist simultaneous with only difference in their proportional

contribution. The non-Darcy flow effect, especially Knudsen flow in this study, has

significant influence on shale gas behavior at low pressure/late-time production.

6. The relationship between apparent permeability and pore pressure can result in

both positive and negative slopes. The negative slope indicates the Knudsen

diffusion effect and slip flow effect have more significant enhancement on

permeability than the reduction from increasing Terzaghi stress. So the true

effective stress coefficient may be calculated separately within different pressure

regions < ~500 psi and > ~500 psi. CO2 apparent permeability can be higher than

He at pressure higher than 1000 psi, which may result from limited shale adsorption

capacity, and thus limited sorption-induced matrix swelling effect when pressure

keeps increasing. However, on the perspective of the analysis based on effective

stress law didn’t provide obvious evidence on the difference between helium and

CO2 permeability evolution.

7. The proposed apparent permeability model under both uniaxial strain condition and

hydrostatic condition potentially have good capacity on unconventional reservoir

rock permeability prediction. Since permeability can vary significantly with

Page 218: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

203

effective stress at high pressure, stress-strain type boundary conditions is critical in

modeling unconventional rock permeability evolution. The hydrostatic model

showed good fit with anthracite permeability data collected under constant stress

condition. By varying pore size, which results from the change of effective stress

and sorption-induced strain, we may obtain the correct fit of permeability data from

nanotube flow-based model. This explains that knowing rock structure change is

critical for determine the promising apparent permeability model with multiple

flow mechanisms. The prediction of shale permeability by this proposed model is

promising based on good data validation results. But further efforts are needed on

solid shale core permeability data measurements under correctly stressed condition.

Page 219: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

204

Chapter 8 Future Work

The potential research objective can be focused on the fully understanding complex gas

flow with different mechanisms in unconventional reservoirs and the development of

apparent permeability model. The interaction between gas and rock, the co-effect of

different flow regimes, such as Darcy flow, slip flow, transition flow and Knudsen

diffusion, and effective stress variation have significant impact on gas permeability during

production. It is necessary to refine the multi-mechanistic approach and figure out the

complicated relationships with simpler solutions. Based on the knowledge gained in this

study, the following topics of research should be pursued further:

1. Laboratory permeability measurement approach should be further refined. Now the

adsorption effect can be only defined in the pore volume parameter and the

calculation process is not optimized enough for large quantity of permeability data.

An approach that include the soption-induced roc matrix swelling/shrinkage effect

or convert sorption to a dynamic process that influence the gas permeability is

needed.

2. CO2 can reach supercritical condition during laboratory injection and certainly it

influences the measured permeability. So a method to quantify this effect is needed.

3. Whether effective stress coefficient can have multiple values for the same sample

same pore fluid need to be further justified. This is critical when a permeability

evolution analysis based on effective stress law is needed. The prediction of

permeability can be totally different due to the selection of effective stress

coefficient.

Page 220: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

205

4. The baseline for differentiate the flow regimes need to be built. Although some

flow regimes, such as slip flow and transition flow, can be expressed in a

combination of Darcy flow and Knudsen diffusion, they may have their own

indispensable flow mechanisms. If so, we should evaluate and properly weighted

them respectively.

5. A better method to couple flow model with fracture-based permeability model may

be developed, since it is still uncertain that the direct combination of flow model

and mechanical model is the best way to build the permeability model. If less

variables can be involved in this analytical model, it will be more generic and easily

used for other academic purpose or industrial application.

Page 221: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

206

References

Aminian, K., & Ameri, S. (2004). Type Curves for Coalbed Methane Production

Prediction. 2004 SPE Eastern Regional Meeting. http://doi.org/10.2523/91482-MS

Amyx, J. W., Bass, D. M. J., & Whiting, R. L. (1960). Petroleum Reservoir Engineering.

An, F. H., Cheng, Y. P., Wu, D. M., & Wang, L. (2013). The effect of small micropores

on methane adsorption of coals from Northern China. Adsorption, 19(1), 83–90.

http://doi.org/10.1007/s10450-012-9421-3

Brace, W. F., Walsh, J. B., & Frangos, W. T. (1968). Permeability of granite under high

pressure. Journal of Geophysical Research, 73(6).

Breckels, I. M., & Eekelen, H. a. M. (1982). Relationship Between Horizontal Stress and

Depth in Sedimentary Basins. Journal of Petroleum Technology, 34(9), 2191–2199.

http://doi.org/10.2118/10336-PA

Busch, A., Gensterblum, Y., Krooss, B. M., & Littke, R. (2004). Methane and carbon

dioxide adsorption–diffusion experiments on coal: upscaling and modeling.

International Journal of Coal Geology, 60(2–4), 151–168.

http://doi.org/10.1016/j.coal.2004.05.002

Cai, Y., Liu, D., Pan, Z., Yao, Y., Li, J., & Qiu, Y. (2013). Pore structure and its impact

on CH4 adsorption capacity and flow capability of bituminous and subbituminous

coals from Northeast China. Fuel, 103, 258–268.

http://doi.org/10.1016/j.fuel.2012.06.055

Choi, J.-G., Do, D. D., & Do, H. D. (2001). Surface Diffusion of Adsorbed Molecules in

Porous Media: Monolayer, Multilayer, and Capillary Condensation Regimes.

Industrial & Engineering Chemistry Research, 40(19), 4005–4031.

http://doi.org/10.1021/ie010195z

Civan, F., Rai, C., & Sondergeld, C. (2011). Shale-gas permeability and diffusivity

inferred by improved formulation of relevant retention and transport mechanisms.

Transport in Porous Media, 86(3), 925–944. http://doi.org/10.1007/s11242-010-

9665-x

Clarkson, C., & Bustin, R. (1999). 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, di, 1333–1344.

Clarkson, C., Pan, Z., Palmer, I., & Harpalani, S. (2010). Predicting sorption-induced

strain and permeability increase with depletion for coalbed-methane reservoirs. SPE

Journal, (March).

Clarkson, C. R., & Bustin, R. M. (1999). The effect of pore structure and gas pressure

upon the transport properties of coal: a laboratory and modeling study. 2. Adsorption

rate modeling. Fuel, 78(11), 1345–1362. http://doi.org/10.1016/S0016-

2361(99)00056-3

Page 222: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

207

Clarkson, C. R., Solano, N., & Bustin, R. (2013). Pore structure characterization of North

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

intrusion. Fuel, 103, 606–616. http://doi.org/10.1016/j.fuel.2012.06.119

Coal Age - DEC 2014. (2014).

Coppens, M. (1999). The effect of fractal surface roughness on diffusion and reaction in

porous catalysts – from fundamentals to practical applications. Catalysis Today,

53(2), 225–243. http://doi.org/10.1016/S0920-5861(99)00118-2

Coppens, M. O., & Dammers, A. J. (2006). Effects of heterogeneity on diffusion in

nanopores - From inorganic materials to protein crystals and ion channels. Fluid

Phase Equilibria, 241(1–2), 308–316. http://doi.org/10.1016/j.fluid.2005.12.039

Crank, J. (1957). The mathematics of diffusion.

Cui, X., Bustin, A., & Bustin, R. (2009). Measurements of gas permeability and

diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 40479.

Cui, X., Bustin, A. M. M., & Bustin, R. M. (2009). Measurements of gas permeability

and diffusivity of tight reservoir rocks: different approaches and their applications.

Geofluids, 9(3), 208–223. http://doi.org/10.1111/j.1468-8123.2009.00244.x

Cui, X., & Bustin, R. M. (2005). Volumetric strain associated with methane desorption

and its impact on coalbed gas production from deep coal seams. AAPG Bulletin,

89(9), 1181–1202. http://doi.org/10.1306/05110504114

Cui, X., Bustin, R. M., & Dipple, G. (2004). Selective transport of CO2, CH4, and N2 in

coals: insights from modeling of experimental gas adsorption data. Fuel, 83(3), 293–

303. http://doi.org/10.1016/j.fuel.2003.09.001

Darabi, H., Ettehad, a., Javadpour, F., & Sepehrnoori, K. (2012). Gas flow in ultra-tight

shale strata. Journal of Fluid Mechanics, 710, 641–658.

http://doi.org/10.1017/jfm.2012.424

Dicker, A. I., & Smits, R. M. (1988). A practical approach for determining permeability

from laboratory pressure-pulse decay measurements. In SPE International Meeting

on Petroleum Engineering, Tianjin, China,1-4 Nov,1998. Tianjin, China.

Do, D. D. (1998). Adsorption Analysis: Equilibria and Kinetics (Vol. 2). Imperial

College Press. http://doi.org/10.1142/9781860943829

Drazer, G., Chertcoff, R., Bruno, L., & Rosen, M. (1999). Concentration dependence of

diffusion–adsorption rate in activated carbon. Chemical Engineering Science, 54,

4285–4291.

Dutta, A. (2009). Multicomponent Gas Diffusion and Adsorption in Coals for Enhanced

Methane Recovery.

EIA. (2015). Marcellus Region Drilling Productivity Report, (March), 2015.

Ertekin, T., King, G., & Schwerer, F. (1986). Dynamic Gas Slippage: A Unique Dual-

Mechanism Approach to the Flow of Gas in Tight Formations. SPE Formation

Page 223: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

208

Evaluation. http://doi.org/10.2118/12045-PA

Gan, H., Nandi, S. P., & Walker, P. L. (1972). Nature of the porosity in American coals.

Fuel, 51(4), 272–277. http://doi.org/10.1016/0016-2361(72)90003-8

Geertsma, J. (1966). Problems of rock mechanics in petroleum production engineering.

1st ISRM Congress, 3(59), 585–594.

Gensterblum, Y., Ghanizadeh, A., & Krooss, B. M. (2014). Gas permeability

measurements on Australian subbituminous coals: Fluid dynamic and poroelastic

aspects. Journal of Natural Gas Science and Engineering, 19, 202–214.

http://doi.org/10.1016/j.jngse.2014.04.016

Halliburton Company. (2007). Chapter 04 - Reservoir Analysis. Coalbed Methane

Princples and Practices, (June 2007), 191–282.

Harpalani, S., & Chen, G. (1997). Influence of gas production induced volumetric strain

on permeability of coal. Geotechnical and Geological Engineering, 15(4), 303–325.

http://doi.org/10.1007/BF00880711

Harpalani, S., & Mitra, A. (2009). Impact of CO2 Injection on Flow Behavior of Coalbed

Methane Reservoirs. Transport in Porous Media, 82(1), 141–156.

http://doi.org/10.1007/s11242-009-9475-1

Harpalani, S., Prusty, B., & Dutta, P. (2006). Methane/CO2 sorption modeling for

coalbed methane production and CO2 sequestration. Energy & Fuels, 20(4), 1591–

1599.

Hartman, R. (2008). Recent advances in the analytical methods used for shale gas

reservoir gas-in-place assessment. In AAPG Annual Convention, San Antonio,

Texas, 20-23 Apr, 2008 (Vol. 40317).

He, W., Lv, W., & Dickerson, J. (2014). Gas Diffusion Mechanisms and Models. In Gas

Transport in Solid Oxide Fuel Cells (pp. 9–17). http://doi.org/10.1007/978-3-319-

09737-4

Heller, R., Vermylen, J., & Zoback, M. (2014). Experimental investigation of matrix

permeability of gas shales. AAPG Bulletin, 98(5), 975–995.

http://doi.org/10.1306/09231313023

Hildenbrand, a., Krooss, B. M., Busch, a., & Gaschnitz, R. (2006). Evolution of

methane sorption capacity of coal seams as a function of burial history — a case

study from the Campine Basin, NE Belgium. International Journal of Coal Geology,

66(3), 179–203. http://doi.org/10.1016/j.coal.2005.07.006

Hsieh, P. a., Tracy, J. V., Neuzil, C. E., Bredehoeft, J. D., & Silliman, S. E. (1981). A

transient laboratory method for determining the hydraulic properties of “tight”

rocks—I. Theory. International Journal of Rock Mechanics and Mining Sciences &

Geomechanics Abstracts, 18(3), 245–252. http://doi.org/10.1016/0148-

9062(81)90979-7

Hubbert, M. K. (1957). Darcy’S Law and the Field Equations of the Flow of

Underground Fluids. International Association of Scientific Hydrology. Bulletin,

Page 224: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

209

2(1), 23–59. http://doi.org/10.1080/02626665709493062

Izadi, G., & Elsworth, D. (2013). The effects of thermal stress and fluid pressure on

induced seismicity during stimulation to production within fractured reservoirs.

Terra Nova, 25(5), 374–380. http://doi.org/10.1111/ter.12046

Izadi, G., Wang, S., Elsworth, D., Liu, J., Wu, Y., & Pone, D. (2011). Permeability

evolution of fluid-infiltrated coal containing discrete fractures. International Journal

of Coal Geology, 85(2), 202–211. http://doi.org/10.1016/j.coal.2010.10.006

Javadpour, F. (2009). Nanopores and apparent permeability of gas flow in mudrocks

(shales and siltstone). Journal of Canadian Petroleum Technology, 48, 1–6.

Javadpour, F., Fisher, D., & Unsworth, M. (2007). Nanoscale gas flow in shale gas

sediments. Journal of Canadian Petroleum Technology, 46(10).

http://doi.org/10.2118/07-10-06

Jones, S. (1997). A technique for faster pulse-decay permeability measurements in tight

rocks. SPE Formation Evaluation, (November 1994), 25–28.

Kamath, J., Boyer, R., & Nakagawa, F. (1992). Characterization of core-scale

heterogeneities using laboratory pressure transients. SPE Formation Evaluation,

(September), 219–227.

Karniadakis, G., Beskok, A., & Narayan, A. (2005). Microflows and Nanoflows.

http://doi.org/10.1007/0-387-28676-4

Klinkenberg, L. J. (1941). The permeability of porous media to liquids and gases.

Drilling and Production Practice, (2), 57–73.

http://doi.org/10.5510/OGP20120200114

Kuila, U., Prasad, M., & Kazemi, H. (2013). Assessing Knudsen flow in gas-flow models

of shale reservoirs (pp. 23–28).

Kumar, A. (2007). Methane Diffusion Characteristics of Illinois Coals.

Kumar, H., Elsworth, D., Liu, J., Pone, D., & Mathews, J. P. (2012). Optimizing

enhanced coalbed methane recovery for unhindered production and CO2 injectivity.

International Journal of Greenhouse Gas Control, 11, 86–97.

http://doi.org/10.1016/j.ijggc.2012.07.028

Kumar, H., Elsworth, D., Mathews, J. P., Liu, J., & Pone, D. (2014). Effect of CO2

injection on heterogeneously permeable coalbed reservoirs. Fuel, 135, 509–521.

http://doi.org/10.1016/j.fuel.2014.07.002

Kumar, H., Elsworth, D., Mathews, J. P., & Marone, C. (2015). Permeability evolution in

sorbing media: analogies between organic-rich shale and coal. Geofluids.

http://doi.org/10.1111/gfl.12135

Kwon, O., Kronenberg, A. K., Gangi, A. F., & Johnson, B. (2001). Permeability of

Wilcox shale and its effective pressure law. Journal of Geophysical Research, 106,

19339. http://doi.org/10.1029/2001JB000273

Langmuir, I. (1918). The adsorption of gases on plane surfaces of glass, mica and

Page 225: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

210

platinum. Journal of the American Chemical Society, 345(1914).

Levine, J. (1996). Model study of the influence of matrix shrinkage on absolute

permeability of coal bed reservoirs. Geological Society, London, Special

Publications.

Li, J., Liu, D., Lu, S., Yao, Y., & Xue, H. (2015). Evaluation and modeling of the CO2

permeability variation by coupling effective pore size evolution in anthracite coal.

Energy & Fuels. http://doi.org/10.1021/ef502674y

Li, J., Liu, D., Yao, Y., Cai, Y., & Chen, Y. (2013). Evaluation and modeling of gas

permeability changes in anthracite coals. Fuel, 111, 606–612.

http://doi.org/10.1016/j.fuel.2013.03.063

Li, Y., Li, X., Shi, J., Wang, H., Wu, L., Teng, S., & Corp, S. (2014). A Nano-Pore Scale

Gas Flow Model for Shale Gas Reservoir Flow characteristic of shale gas reservoir

in various pole scales and molecule scales, (June), 9–11.

Liu, J., Chen, Z., Elsworth, D., Miao, X., & Mao, X. (2010). Linking gas-sorption

induced changes in coal permeability to directional strains through a modulus

reduction ratio. International Journal of Coal Geology, 83(1), 21–30.

http://doi.org/10.1016/j.coal.2010.04.006

Liu, S. (2012). Estimation of different coal compressibilities of coalbed methane

reservoirs under replicated in situ condition.

Liu, S., & Harpalani, S. (2013a). A new theoretical approach to model sorption-induced

coal shrinkage or swelling. AAPG Bulletin, 97(7), 1033–1049.

http://doi.org/10.1306/12181212061

Liu, S., & Harpalani, S. (2013b). Permeability prediction of coalbed methane reservoirs

during primary depletion. International Journal of Coal Geology, 113, 1–10.

http://doi.org/10.1016/j.coal.2013.03.010

Liu, S., & Harpalani, S. (2014a). Compressibility of sorptive porous media: Part 1.

Background and theory. AAPG Bulletin, c, 1–25.

Liu, S., & Harpalani, S. (2014b). Compressibility of Sorptive Porous Media: Part I –

Background and Theory. AAPG Bulletin, 98(9), 1761–1772.

Liu, S., & Harpalani, S. (2014c). Compressibility of Sorptive Porous Media: Part II –

Experimental Study on Coal. AAPG Bulletin, 98(9), 1773–1788.

Liu, S., & Harpalani, S. (2014d). Determination of the effective stress law for

deformation in coalbed methane reservoirs. Rock Mechanics and Rock Engineering,

47(5), 1809–1820. http://doi.org/10.1007/s00603-013-0492-6

Liu, S., & Harpalani, S. (2014e). Evaluation of in situ stress changes with gas depletion

of coalbed methane reservoirs. Journal of Geophysical Research: Solid Earth,

119(8), 6263–6276. http://doi.org/10.1002/2014JB011228.

Liu, S., Harpalani, S., & Pillalamarry, M. (2012). Laboratory measurement and modeling

of coal permeability with continued methane production: Part 2 – Modeling results.

Fuel, 94, 117–124. http://doi.org/10.1016/j.fuel.2011.10.053

Page 226: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

211

Lorenz, J. C., Teufel, L. W., & Warpinski, N. R. (1991). Regional fractures; I, A

mechanism for the formation of regional fractures at depth in flat-lying reservoirs.

AAPG Bulletin , 75(11), 1714–1737.

Luffel, D., Hopkins, C., & Jr, P. S. (1993). Matrix permeability measurement of gas

productive shales. In SPE Annual Technical Conference and Exhibition, Houston,

Texas, 2-6 Oct, 1993.

Ma, Q., Harpalani, S., & Liu, S. (2011). A simplified permeability model for coalbed

methane reservoirs based on matchstick strain and constant volume theory.

International Journal of Coal Geology, 85(1), 43–48.

http://doi.org/10.1016/j.coal.2010.09.007

Malkovsky, V. I., Zharikov, a. V., & Shmonov, V. M. (2009). New methods for

measuring the permeability of rock samples for a single-phase fluid. Izvestiya,

Physics of the Solid Earth, 45(2), 89–100.

http://doi.org/10.1134/S1069351309020013

Markowski, A. K. (2014). PENNSYLVANIA COALBED METHANE UPDATE, 80353.

Markowski, A., Resources, N., & Virginia, W. (2014). Coalbed methane in

Pennsylvania.

Mavor, M. J., Owen, L. B., & Pratt, T. J. (1990). Measurement and Evaluation of Coal

Sorption Isotherm Data. Proceedings of SPE Annual Technical Conference and

Exhibition, (1). http://doi.org/10.2523/20728-MS

Mazumder, S., & Wolf, K. (2008). Differential swelling and permeability change of coal

in response to CO 2 injection for ECBM. International Journal of Coal Geology.

Mckernan, R. E., Rutter, E. H., Mecklenburgh, J., & Taylor, K. G. (2014). Influence of

Effective Pressure on Mudstone Matrix Permeability: Implications for Shale Gas

Production. In SPE/EAGE European Unconventional Conference and Exhibition,

Vienna, Austria, 25-27 Feb, 2014 (p. 13). Society of Petroleum Engineers.

http://doi.org/10.2118/167762-MS

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1–Laboratory results. Fuel,

94, 110–116.

Mitra, A., Harpalani, S., & Liu, S. (2012). Laboratory measurement and modeling of coal

permeability with continued methane production: Part 1 – Laboratory results. Fuel,

94, 110–116. http://doi.org/10.1016/j.fuel.2011.10.052

Okotie, V. U., & Moore, R. L. (2011). Well-production challenges and solutions in a

mature, very-low-pressure coalbed-methane reservoir. SPE Production and

Operations, 26(2), 149–161. http://doi.org/10.2118/107705-MS.

Palmer, I., & Mansoori, J. (1998). How permeability depends on stress and pore pressure

in coalbeds: a new model. SPE Reservoir Evaluation and Engineering, SPE

52607(December), 539–544.

Palmer, S. I., & Mansoori, J. (1998). How permeability depends on stress and pore

Page 227: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

212

pressure in coalbeds: a new model. Soc Petrol Eng Res Eng, (December), 539–544.

Pan, Z., & Connell, L. D. (2007). A theoretical model for gas adsorption-induced coal

swelling. International Journal of Coal Geology, 69(4), 243–252.

http://doi.org/10.1016/j.coal.2006.04.006

Pillalamarry, M., Harpalani, S., & Liu, S. (2011). Gas diffusion behavior of coal and its

impact on production from coalbed methane reservoirs. International Journal of

Coal Geology, 86(4), 342–348. http://doi.org/10.1016/j.coal.2011.03.007

Randolph, P. L., Soeder.D.J., & Chowdiah, P. (1984). Porosity and Permeability of Tight

Sands. In SPE/DOE/GRI Unconventional Gas Recovery Symposium held in

Pittsburgh, PA, May 13-15. Dallas.

Reif.F. (1965). Fundamentals of statistical and thermal physics. Retrieved from

http://catalog.hathitrust.org/Record/000145539/Cite

Rexer, T. F. T., Benham, M. J., Aplin, A. C., & Thomas, K. M. (2013). Methane

Adsorption on Shale under Simulated Geological Temperature and Pressure

Conditions. Energy & Fuels, 27(6), 3099–3109. http://doi.org/10.1021/ef400381v

Rodrigues, C. F., & Lemos De Sousa, M. J. (2002). The measurement of coal porosity

with different gases. International Journal of Coal Geology, 48(3–4), 245–251.

http://doi.org/10.1016/S0166-5162(01)00061-1

Roy, S., Raju, R., Chuang, H. F., Cruden, B. a., & Meyyappan, M. (2003). Modeling gas

flow through microchannels and nanopores. Journal of Applied Physics, 93(8),

4870–4879. http://doi.org/10.1063/1.1559936

Ruckenstein, E. (1971). Sorption by solids with bidisperse pore structures. Chemical

Engineering …, 26, 1305–1318. Retrieved from

http://www.sciencedirect.com/science/article/pii/0009250971800519

Rydzy, M. B., Patino, J., Elmetni, N., & Appel, M. (2016). Stressed permeability in

shales : effects of matrix compressibility and fractures - a step towards measuring

matrix permeability in fractured shale samples, 1–12.

Sakhaee-Pour, A., & Bryant, S. (2012). Gas Permeability of Shale. SPE Reservoir

Evaluation & Engineering, 15, 401–409.

Seidle, J. P., Jeansonne, M. W., & Erickson, D. J. (1992). Application of matchstick

geometry to stress dependent permeability in coals. SPE Rocky Mountain Regional

Meeting, 433–444. http://doi.org/10.2523/24361-MS

Shabro, V. (2012). Forecasting Gas Production in Organic Shale with the Combined

Numerical Simulation of Gas Diffusion in Kerogen, Langmuir Desorption from

Kerogen Surfaces, and Advection in Nanopores. SPE Annual Technical ….

http://doi.org/10.2118/159250-MS

Shi, J.-Q., Pan, Z., & Durucan, S. (2014). Analytical models for coal permeability

changes during coalbed methane recovery: Model comparison and performance

evaluation. International Journal of Coal Geology, 136, 17–24.

http://doi.org/10.1016/j.coal.2014.10.004

Page 228: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

213

Shi, J., & Durucan, S. (2005). A model for changes in coalbed permeability during

primary and enhanced methane recovery. SPE Reservoir Evaluation & Engineering,

8(4), 291–299.

Shi, J. Q., & Durucan, S. (2003). A bidisperse pore diffusion model for methane

displacement desorption in coal by CO 2 injection☆. Fuel, 82(10), 1219–1229.

http://doi.org/10.1016/S0016-2361(03)00010-3

Shi, J. Q., & Durucan, S. (2004). Drawdown Induced Changes in Permeability of

Coalbeds: A New Interpretation of the Reservoir Response to Primary Recovery.

Transport in Porous Media, 56(1), 1–16.

http://doi.org/10.1023/B:TIPM.0000018398.19928.5a

Shi, J. Q., & Durucan, S. (2014). Modelling laboratory horizontal stress and coal

permeability data using S&D permeability model. International Journal of Coal

Geology, 131, 172–176. http://doi.org/10.1016/j.coal.2014.06.014

Singh, H., & Javadpour, F. (2013). Nonempirical Apparent Permeability of Shale. In

Unconventional Resources Technology Conference (Vol. 2, p. 16). Denvor,

Colorado.

Singh, H., & Javadpour, F. (2015). Langmuir slip-Langmuir sorption permeability model

of shale. FUEL, (October). http://doi.org/10.1016/j.fuel.2015.09.073

Smith, D., & Williams, F. (1984). Diffusional effects in the recovery of methane from

coalbeds. SPEJ, Soc. Pet. Eng. J.;(United States), (October), 529–535.

Soeder, D. (1988). Porosity and permeability of eastern Devonian gas shale. SPE

Formation Evaluation, 3, 116–124.

Sudibandriyo, M., Pan, Z., Fitzgerald, J. E., Jr, R. L. R., & Gasem, K. A. M. (2003).

Adsorption of Methane, Nitrogen, Carbon Dioxide, and Their Binary Mixtures on

Dry Activated Carbon at 318.2 K and Pressures up to 13.6 MPa. Lan, 19(8), 5323–

5331.

Suggested Methods for Uniaxial‐Strain Compressibility Testing for Reservoir

Geomechanics. (2015), c(April), 1–29.

Swami, V. (2012). Shale Gas Reservoir Modeling: From Nanopores to Laboratory. SPE

Annual Technical Conference and Exhibition, (October), 8–10.

Swami, V., Clarkson, C., & Settari, A. (2012). Non-Darcy Flow in Shale Nanopores: Do

We Have a Final Answer? SPE Canadian Unconventional …, (November), 1–17.

Swami, V., & Settari, A. (2012). A pore scale gas flow model for shale gas reservoir. In

SPE Americas Unconventional Resources Conference held in Pittsburgh (pp. 5–7).

Terzaghi, K. (1943). Theoretical Soil Mechanics. John Wiley & Sons, Inc.

http://doi.org/10.1680/geot.1964.14.1.1

Teufel, L. W., Rhett, D. W., & Farrell, H. E. (1991). Effect of Reservoir Depletion And

Pore Pressure Drawdown On In Situ Stress And Deformation In the Ekofisk Field,

North Sea. The 32nd U.S. Symposium on Rock Mechanics (USRMS), 63–72.

Page 229: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

214

U S Energy Information Administration. (2011). Emissions of Greenhouse Gases in the

United States 2009, (March).

Walls, J., Diaz, E., & Cavanaugh, T. (2012). Shale Reservoir Properties from Digital

Rock Physics. Proceedings of SPE/EAGE European Unconventional Resources

Conference and Exhibition, 1–9. http://doi.org/10.2118/152752-MS

Wang, G., Ren, T., Wang, K., & Zhou, A. (2014). Improved apparent permeability

models of gas flow in coal with Klinkenberg effect. Fuel, 128, 53–61.

http://doi.org/10.1016/j.fuel.2014.02.066

Wang, S., Elsworth, D., & Liu, J. (2011). Permeability evolution in fractured coal: the

roles of fracture geometry and water-content. International Journal of Coal

Geology, 87(1), 13–25. http://doi.org/10.1016/j.coal.2011.04.009

Wang, S., Elsworth, D., & Liu, J. (2012). A mechanistic model for permeability evolution

in fractured sorbing media. Journal of Geophysical Research, 117(B6).

Wang, S., Elsworth, D., & Liu, J. (2012a). A mechanistic model for permeability

evolution in fractured sorbing media. Journal of Geophysical Research: Solid Earth

( …, 117(B6), B06205. http://doi.org/10.1029/2011JB008855

Wang, S., Elsworth, D., & Liu, J. (2012b). Permeability evolution during progressive

deformation of intact coal and implications for instability in underground coal

seams. International Journal of Rock Mechanics and Mining Sciences, 1–12.

Wang, Y., & Liu, S. (2016). Estimation of Pressure-Dependent Diffusive Permeability of

Coal Using Methane Diffusion Coefficient: Laboratory Measurements and

Modeling. Energy and Fuels, 30(11), 8968–8976.

http://doi.org/10.1021/acs.energyfuels.6b01480

Wang, Y., Liu, S., & Elsworth, D. (2015). Laboratory investigations of gas flow

behaviors in tight anthracite and evaluation of different pulse-decay methods on

permeability estimation. International Journal of Coal Geology, 149, 118–128.

http://doi.org/10.1016/j.coal.2015.07.009

Wu, K., Li, X., Wang, C., Chen, Z., & Yu, W. (2015). A Model for Gas Transport in

Microfractures of Shale and Tight Gas Reservoirs. AIChE Journal, 61(6).

http://doi.org/10.1002/aic

Wu, K., Li, X., Wang, C., Yu, W., & Chen, Z. (2014). Apparent Permeability for Gas

Flow in Shale Reservoirs Coupling Effects of Gas Diffusion and Desorption.

Proceedings of the 2nd Unconventional Resources Technology Conference, 1–18.

http://doi.org/10.15530/urtec-2014-1921039

Xu, H., Tang, D., Zhao, J., Li, S., & Tao, S. (2015). A new laboratory method for

accurate measurement of the methane diffusion coefficient and its influencing

factors in the coal matrix. Fuel, 158, 239–247.

http://doi.org/10.1016/j.fuel.2015.05.046

Yi, J., Akkutlu, I., & Deutsch, C. (2008). Gas Transport in Bidisperse Coal Particles:

Investigation for an Effective Diffusion Coefficient in Coalbeds. Journal of

Page 230: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

215

Canadian Petroleum Technology, 47(10). http://doi.org/10.2118/08-10-20

Yin, G., Jiang, C., Wang, J. G., & Xu, J. (2013). Combined Effect of Stress, Pore

Pressure and Temperature on Methane Permeability in Anthracite Coal: An

Experimental Study. Transport in Porous Media, 100(1), 1–16.

http://doi.org/10.1007/s11242-013-0202-6

Yuan, W., Pan, Z., Li, X., Yang, Y., Zhao, C., Connell, L. D., … He, J. (2014).

Experimental study and modelling of methane adsorption and diffusion in shale.

Fuel, 117, 509–519. http://doi.org/10.1016/j.fuel.2013.09.046

Zhang, T., Ellis, G. S., Ruppel, S. C., Milliken, K., & Yang, R. (2012). Effect of organic-

matter type and thermal maturity on methane adsorption in shale-gas systems.

Organic Geochemistry, 47, 120–131.

http://doi.org/10.1016/j.orggeochem.2012.03.012

Zoback, M. D., & Byerlee, J. D. (1975). Permeability and Effective Stress. Aapg Bulletin-

American Association of Petroleum Geologists, 59(1), 154–158.

http://doi.org/10.1306/83D91C40-16C7-11D7-8645000102C1865D

Page 231: LABORATORY ESTIMATION AND MODELING OF APPARENT ...

VITA

Yi Wang

EDUCATION

The Pennsylvania State University, University Park, PA, 2010-2012

Master of Science, Petroleum & Natural Gas Engineering

China University of Petroleum, Beijing, China, 2006-2010

Bachelor of Science, Petroleum & Natural Gas Engineering

PUBLICATIONS

Yi Wang and Shimin Liu. 2016. Estimation of Pressure-dependent Diffusive Permeability of Coal Using

Methane Diffusion Coefficient: Laboratory Measurements and Modeling. Energy & Fuels. 20 pp.

Shimin Liu, Yi Wang and Satya Harpalani. 2016. Anisotropy characteristics of coal shrinkage/swelling

and its impact on coal permeability evolution with CO2 injection. Greenhouse Gases: Science and

Technology. 18 pp.

Shun Liang, Derek Elsworth, Xuehua Li, Xuehai Fu, Dong Yang, Qiangling Yao and Yi Wang. 2015.

Dynamic impacts on the survivability of shale gas wells piercing longwall panels. Journal of

Natural Gas Science and Engineering. 16pp.

Rui Zhang, Shimin Liu, Jitendra Bahadur, Derek Elsworth, Yuri Melnichenko, Lilin He and Yi Wang.

2015. Estimation and modeling of coal pore accessibility using small angle neutron scattering. Fuel.

9pp.

Yi Wang, Shimin Liu and Derek Elsworth. 2015. Laboratory investigations of gas flow behaviors in

tight anthracite and evaluation of different pulse-decay methods on permeability estimation.

International Journal of Coal Geology. 10 pp.

Yi Wang, Robert Watson, Jamal Rostami, John Yilin Wang, Mark Limbruner and Zhong He. 2013.

Study of borehole stability of Marcellus shale wells in longwall mining areas. Journal of Petroleum

Exploration and Production Technology. 13 pp.

Yi Wang and Shimin Liu. 2015. Shale core apparent permeability characterization with CO2 injection

under stress-controlled condition. The 14th Annual Carbon Capture, Utilization & Storage

Conference, Pittsburgh, Pennsylvania. 12pp.

Shimin Liu, Satya Harpalani and Yi Wang. 2013. Deformation characteristics of coal and its impact on

permeability during CO2 sequestration. International Conference on Coal Science & Technology.

13 pp.