Você está na página 1de 141

NEW COUPLING CONSIDERATIONS BETWEEN MATRIX AND MULTISCALE

NATURAL FRACTURES IN UNCONVENTIONAL


RESOURCE RESERVOIRS






by
Osman Apaydin
ii

A dissertation submitted to the Faculty and the Board of Trustees of the Colorado School of
Mines in partial fulfillment of the requirements for the degree of Doctor of Philosophy
(Petroleum Engineering).


Golden, Colorado
Date: ______________________




Signed: ______________________________
Osman Apaydin

Approved: ____________________________
Dr. Erdal Ozkan
Professor
Dissertation Advisor



Golden, Colorado
Date: ________________________



Signed: ______________________________
Dr. Ramona M. Graves
Professor and Head
Department of Petroleum Engineering




iii

ABSTRACT
This dissertation presents two analytical models to describe gas flow in shale matrix and
gas transfer from matrix to fracture network in fractured shalegas reservoirs. The first model
considers the effect of microfractures on shale while the second model incorporates slip flow into
matrix-flow models and pressure-dependent permeability in reservoir macrofractures. These
models can be extended to liquid-rich reservoirs.
In the first model, the effects of matrix microfractures on effective matrix permeability of
a dual-porosity medium are examined. An analytical model is presented with composite matrix
blocks consisting of a core where unconnected microfractures do not considerably contribute to
flow capacity and a surface layer where the microfractures connected to the matrix surface cause
a stimulation effect. The composite matrix flow is coupled with the flow in a network of
macrofractures. The dissertation investigates the effect of matrix-surface stimulation and
demonstrates improved fluid transfer from the matrix medium to the fracture network due to
matrix microfractures. It is shown that matrix microfractures accelerate production by providing
earlier and more effective contribution of matrix into flow rates. This contribution of the matrix
due to microfractures cannot be simulated by enhanced matrix permeability as the microfractured
surface-layer of the matrix causes different flow characteristics than a homogeneous (unfractured)
matrix.
In the second model, the description of matrix flow is improved by considering diffusive
(Knudsen) flow in nanopores. In this dual-mechanism approach (that accounts for Darcy and slip
flow), when Darcy flow becomes insignificant due to nanodarcy matrix permeability, diffusive
flow takes over and contributes, substantially, to the transfer of fluids from matrix to fracture
network. Furthermore, stress-dependent permeability in the fracture network is considered.
Therefore, incorporating Darcy and diffusive flows in the matrix and stress-dependent
permeability in the fractures, a dual-mechanism dual-porosity naturally fractured reservoir
formulation is developed and a new transfer function for fractured shale-gas reservoirs is derived.
The dual-mechanism dual-porosity formulation presented in this dissertation can be used for
numerical or analytical modeling purposes. The new formulation of matrix to fracture fluid
transfer with an analytical model is used to demonstrate the differences from the conventional
formulation.

iv


TABLE OF CONTENTS
ABSTRACT .................................................................................................................................. iii
LIST OF FIGURES ....................................................................................................................... vii
LIST OF TABLES ........................................................................................................................... x
ACKNOWLEDGEMENT .............................................................................................................. xi
CHAPTER 1 INTRODUCTION ..................................................................................................... 1
1.1 Organization of the Dissertation ........................................................................................... 2
1.2 Motivation of Research ......................................................................................................... 2
1.3 Hypotheses ............................................................................................................................ 4
1.4 Objectives .............................................................................................................................. 5
1.5 Phases of the Research .......................................................................................................... 5
1.6 Method of Study .................................................................................................................... 6
CHAPTER 2 LITERATURE REVIEW .......................................................................................... 7
CHAPTER 3 MICROFRACTURES, DIFFUSIVE FLOW, AND STRESS-DEPENDENT
NATURAL FRACTURE PERMEABILITY ......................................................... 11
3.1 Microfractures in Source Rocks .......................................................................................... 11
3.2 Knudsen Diffusion .............................................................................................................. 14
3.3 Stress-Dependent Fracture Permeability Reduction ........................................................... 23
CHAPTER 4 ANALYTICAL FLOW MODELS .......................................................................... 27
4.1 General Assumptions, Definitions of Dimensionless Variables and Model Parameters .... 27
4.2 Dual-Porosity Model with Matrix, Microfracture, and Macrofracture Media .................... 28
4.2.1 Flow in a Microfractured Spherical Matrix Block .................................................... 29
4.2.3 Flow in the Macrofracture Network .......................................................................... 32
4.2.4 Verification of the Solution ....................................................................................... 33
4.2.5 Extension to Multiple Fractured Horizontal Wells with Stimulated
Reservoir Volume ..................................................................................................... 35
v

4.3 Dual-Porosity Model with Knudsen Diffusion (Slip Flow) in Matrix and
Stress-Dependent Fracture Permeability Reduction ........................................................... 36
4.3.1 Introduction of Knudsen Diffusion (Slip Flow) ........................................................ 37
4.3.2 Introduction of Stress-Dependent Natural-Fracture Permeability ............................. 38
4.3.3 Dual-Mechanism Dual-Porosity Reservoir Model and Transfer Function ............... 41
CHAPTER 5 DISCUSSION OF RESULTS ................................................................................. 46
5.1 Dual-Porosity Model with Matrix, Microfracture and Macrofracture Media ..................... 46
5.2 Dual-Porosity Model with Knudsen Diffusion (Slip Flow) in Matrix and
Stress-Dependent Fracture Permeability ............................................................................ 59
5.2.1 Comparison of the Dual-Porosity Formulations with Spherical Matrix Blocks
and Slabs ................................................................................................................... 59
5.2.2 Production from Source Matrix ................................................................................. 60
5.2.3 Effect of Slip Flow in Matrix and Stress-Dependent Permeability on the
Performances of Fractured Horizontal Wells ........................................................... 62
CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS .................................................. 67
6.1 Conclusions ......................................................................................................................... 67
6.1.1 Effect of (Matrix-Surface) Microfractures ................................................................ 67
6.1.2 Combined Effect of Slip-Flow and Stress-Dependent Permeability ......................... 68
6.2 Recommendations for Future Work .................................................................................... 69
NOMENCLATURE ...................................................................................................................... 71
REFERENCES CITED .................................................................................................................. 74
APPENDIX A DERIVATION OF THE SOLUTION FOR FLOW IN A
MICROFRACTURED SPHERICAL MATRIX BLOCK .................................... 79
Appendix A.1 Flow in the Matrix Core: ................................................................................... 79
Appendix A.2 Flow in the Matrix Surface Layer: .................................................................... 83
Flow in the Surface-Layer Microfractures: ........................................................................ 83
Flow in the Surface-Layer Matrix: ..................................................................................... 86
Coupled Flow Solution for the Matrix Surface Layer: ....................................................... 89
vi

APPENDIX B DERIVATION OF THE SOLUTION FOR RADIAL FLOW IN THE
MACROFRACTURE NETWORK ...................................................................... 91
APPENDIX C DERIVATION OF DUAL-MECHANISM DUAL-POROSITY MODEL
FOR SHALE-GAS RESERVOIRS ....................................................................... 97
Appendix C.1 Dual-Mechanism Flow in a Spherical Matrix Block ......................................... 97
Appendix C.2 Flow in Fracture Network with Stress-Dependent Permeability ..................... 105
Appendix C.3 Coupling Fracture and Matrix Flows to derive a Dual-Mechanism Dual-
Porosity Transfer Function ............................................................................. 110
APPENDIX D DERIVATION OF DUAL-MECHANISM DUAL-POROSITY MODEL
WITH MICROFRACTURES FOR SHALE-GAS RESERVOIRS .................... 112
Appendix D.1 Flow in the Matrix Core: ................................................................................. 112
Appendix D.2 Flow in the Matrix Surface Layer: .................................................................. 117
Flow in the Surface-Layer Microfractures: ...................................................................... 117
Flow in the Surface-Layer Matrix: ................................................................................... 120
Coupled Flow Solution for the Matrix Surface Layer: ..................................................... 123
Appendix D.3 Radial Dual-Porosity Model with Transient Flow in Spherical Matrix
Blocks: ............................................................................................................ 125


vii

LIST OF FIGURES
Figure 3.1 Volume changes taking place in kerogen and generated hydrocarbons during
increasing thermal maturity (Meissner, 2007). .......................................................... 12
Figure 3.2 Volume changes taking place in oil being cracked to more thermally stable
hydrocarbon species plus residual solid dead oil or coke during increasing
thermal maturity. (Meissner, 2007). ........................................................................... 12
Figure 3.3 (A) Pressure/ depth gradient from normal pressure to overpressure across
hydraulic seal surrounding overpressured rock section. (B) Pressure/ depth
gradient for similar underpressured rock system (Hunt, 1990). ................................. 13
Figure 3.4 Theoretical change in water pressure caused by temperature changeswith
volume remaining constant. The Aquithermal effect (Meissner, 2007). ................ 13
Figure 3.5 (A) and (B): microfractures propagating between kerogen particles (Engelder
et al. 2008). ................................................................................................................. 15
Figure 3.6 Partly mineralized fracture from Marcellus Shale in Pennsylvania (Soeder et al.
2009). .......................................................................................................................... 16
Figure 3.7 Dual-porosity idealizations of naturally fractured reservoirs (Najurieta, 1980). ....... 16
Figure 3.8 Composite matrix model to idealize flow in microfractured shale matrix. ................ 17
Figure 3.9 Composite Sizes of molecules and pore throats in siliciclastic rocks on a
logarithmic scale covering seven orders of magnitude (Modified from Nelson,
2009, by Terri Olson, 2012). ...................................................................................... 18
Figure 3.10 Results of mercury-porosimetery analysis of the Barnett Shale (Modified from
Bowker, 2007b by Bruner et al., 2011). .................................................................. 20
Figure 3.11 Pore-throat diameters calculated from four capillary-pressure sample analyses
of Barnett Shale in the Blakey #1 well (Loucks et al. 2009). ................................... 20
Figure 3.12 Secondary Electron Images showing variations of morphology in organic
matter (Loucks et al. 2009). ...................................................................................... 22
Figure 3.13 Extended Extraction Measurements from a Niobrara Core. .................................... 22
Figure 3.14 Knudsen number regimes and governing equations (Roy et. al. 2003). ................... 24
Figure 3.15 (A) High mean flowing pressure (B) Low mean flowing pressure ........................ 24
Figure 3.16 Flow regimes, pore diameter ranges and permeability ranges for conventional,
tight and shale oil and gas systems. .......................................................................... 24
Figure 3.17 Gas flow mechanisms at different scales (Javadpour et al., 2007). .......................... 25
Figure 3.18 Flow regimes, pore diameter ranges and permeability ranges for conventional,
tight and shale oil and gas systems (Figure from Javadpour et al. 2007). ................ 25
viii

Figure 4.1 Schematic of a radial dual-porosity medium with spherical matrix blocks. .............. 29
Figure 4.2 Spherical matrix block with a composite structure. ................................................... 30
Figure 4.3 Representation of the fractured matrix-surface layer by a system of slabs in
parallel. ....................................................................................................................... 30
Figure 4.4 Geometry of the matrix and fracture slabs used to represent the matrix surface
layer. ........................................................................................................................... 30
Figure 4.5 Trilinear-flow model consisting of a fractured inner reservoir (SRV) and a
homogenous outer reservoir (Ozkan et al., 2009, and Brown et al., 2009). ............... 35
Figure 4.6 Contribution of diffusive flow to the apparent matrix permeability. ......................... 39
Figure 4.7 Error caused by neglecting diffusive flow in matrix for a 0.8 specific gravity gas. .. 39
Figure 4.8 Stress-dependent fracture permeability and the error for neglecting the stress
dependency as a function of pressure. ........................................................................ 40
Figure 4.9 Schematic of the system used to incorporate matrix diffusive flow into transient
dual-porosity model. ................................................................................................... 41
Figure 5.1 Comparison of production performances of a fractured horizontal well in
homogeneous reservoir and dual-porosity reservoirs with and without
microfractures. ............................................................................................................ 49
Figure 5.2 Effect of matrix surface layer thickness (microfracture penetration into matrix)
on production performances of a fractured horizontal well in a dual-porosity
reservoir with microfractures. .................................................................................... 51
Figure 5.3 Effect of matrix permeability on the impact of microfractures on production
performances of a fractured horizontal well. ............................................................. 52
Figure 5.4 Effect of microfracture permeability on production performance of a fractured
horizontal well. ........................................................................................................... 53
Figure 5.5 Effect of microfractures on pseudopressure responses of a fractured horizontal
well. ............................................................................................................................ 54
Figure 5.6 Production history of a fractured, horizontal gas well in the Haynesville Shale
(Wang and Liu, 2011). ............................................................................................... 55
Figure 5.7 Pressure match of the Haynesville shale gas reservoir well with/without
microfracs (Matrix Permeability; 3.2E-06 md). ......................................................... 57
Figure 5.8 Production history of a horizontal gas well in the Barnett Shale (Anderson, 2010). . 57
Figure 5.9 Pressure match of the Barnett shale gas reservoir well with/without microfracs
(Matrix Permeability; 2.9E-04 md). ........................................................................... 59
ix

Figure 5.10 Comparison of pseudopressure and derivative responses for dual-porosity
models with slab and spherical matrix blocks. ......................................................... 61
Figure 5.11 Schematic of a spherical shale-matrix block adjacent to a horizontal well
fracture. ..................................................................................................................... 61
Figure 5.12 Matrix flow rate profiles after 27.3 years of production. ......................................... 62
Figure 5.13 Combined effect of slip flow and stress-dependent natural fracture permeability
on pseudopressure and derivative responses. ........................................................... 63
Figure 5.14 Combined effect of slip flow and stress-dependent natural fracture permeability
on pressure responses. .............................................................................................. 64
Figure 5.15 Effect of stress-dependent fracture permeability on pseudopressure and
derivative responses. ................................................................................................. 65
Figure 5.16 Effect of stress-dependent fracture permeability on pressure responses. ................. 66



x

LIST OF TABLES
Table 5.1 Synthetic Data Used for Results Presented in Figs. 5.1 through 5.5 ............................ 48
Table 5.2 Data used for Matching the Haynesville Shale-Gas Well Performance ....................... 56
Table 5.3 Data used for Matching the Barnett Shale-Gas Well Performance .............................. 58
Table 5.4 Data Used for Results ................................................................................................... 60


xi

ACKNOWLEDGEMENT
I would like to gratefully acknowledge the guidance and advice of my advisor, Dr. Erdal
Ozkan and thank him and his wife, Aysin, for their years of friendship and support in my personal,
professional and school life since my undergraduate years. Without Dr. Ozkans support and his
extensive contributions, this research and dissertation would not have been possible. Additionally,
I would like to thank the dissertation committee members, Dr Kazemi, Dr. Ozbay, Dr.
Miskimmins and Dr. Godesiabois, for their guidance on the research. Special thanks to the
Petroleum Engineering Departments staff, especially Denise Winn-Bower, for their help during
my school years at Colorado School of Mines.
I also would like to thank Schlumberger Data and Consulting Services for supporting me
during initial class work phase of my studies. I would like to express my appreciation of EOG
Resources for both financial and technical support. My last 6 years with EOG Resources has
helped me understand the flow mechanisms in unconventional reservoirs. The work environment
at EOG Resources encouraged me to reach my potential by continuously supporting my technical
and professional improvement. I would also like to acknowledge the Marathon Center of
Excellence for Reservoir Studies (MCERS) at Colorado School of Mines for providing a
cultivating research environment.
I would like to express my sincere gratitude and appreciation to my wife, Tuba, my
daughter, Ekin and my son Ozan. Their full support, sharing, understanding and sacrifices during
the preparation of this dissertation were invaluable. And finally I would like extend my great
respect to my parents and thank them for all the encouragement they provided throughout my life
and that helped me become who I am.
Thank you.
1

CHAPTER 1
INTRODUCTION
This research study is for a Doctor of Philosophy degree that was conducted at the
Marathon Center of Excellence for Reservoir Studies (MCERS) in the Petroleum Engineering
Department of the Colorado School of Mines.
The objective of the research is to improve the description of fluid flow in shale matrix
and fluid transfer from matrix to fracture network in modeling production from fractured shale
gas and liquidsrich reservoirs. In the research, a three-tier approach is used. First, an analytical
model is developed to consider the contribution of matrix microfractures to matrix-to-fracture
fluid transfer. This model represents the matrix blocks as composite media consisting of a core
where unconnected microfractures do not considerably contribute to flow capacity and a surface
layer where the microfractures connected to the matrix surface (resembling wormholes) cause a
stimulation effect. The composite matrix flow is then coupled with the flow in a network of
macrofractures as in the conventional dual-porosity idealizations of fractured media. This model
is used to investigate the contribution of matrix-surface stimulation caused by microfractures to
fluid transfer from the matrix medium to the fracture network and the resulting improvement of
productivity of shale-gas or liquids-rich-reservoir wells.
In the second part of the research, the description of gas flow in shale matrix is improved
by considering diffusive (slip) flow in nanopores. Currently, most modeling approaches for shale-
gas production are based on the dominance of Darcy flow in both natural fractures and matrix. In
the developed dual-mechanism approach, when Darcy flow becomes insignificant due to
nanodarcy matrix permeability, diffusive (slip) flow takes over and contributes, substantially, to
gas flow in shale matrix. To further improve the definition of flow in fractured shale, stress-
dependent permeability in the fracture network is also considered in the model. Therefore,
incorporating Darcy and diffusive flows in the matrix and stress-dependent permeability in the
fractures, a dual-mechanism dual-porosity naturally fractured reservoir formulation is developed
and a new transfer function for fractured shale-gas reservoirs is derived. One of the major
challenges of the derivation of this dual-mechanism dual-porosity model is in the representation
of the boundary condition between matrix and fractures. Incorporation of slip-flow introduces
second-order terms into matrix-flow velocity and gives rise to a mismatch when coupled with the
first-order (Darcy) fracture-flow velocity at the matrix-fracture interface. The dual-mechanism
2

dual-porosity formulation proposed here will be used with an analytical model to demonstrate the
differences from the conventional formulation.
1.1 Organization of the Dissertation
This dissertation explains the motivation, methodology, hypothesis, and the objectives of
the conducted study and presents the results. It is divided into six chapters.
Chapter 1 is the introduction that explains the motivation behind the study and states the
hypotheses of the conducted research along with the objectives and methods of the study.
Chapter 2 presents a review of the literature which is relevant for the conducted research.
Chapter 3 introduces the concepts for non-connected microfractures in source rocks, diffusive
flow at low pressures and tight formations, correlations for fracture deformation and stress-
dependent permeability.
Chapter 4 contains the derivation of the analytical formulation and explains the basic approach as
well as the mathematical issues.
Chapter 5 discusses the results of the research including the impact of incorporation of
microfractures, diffusive flow and stress-dependent permeability in fractures on well productivity
estimations.
Chapter 6 summarizes the conclusions and recommendations of the research with comments on
its potential extensions.
1.2 Motivation of Research
In recent years, gas and oil production from resource rocks has become an important
portion of our hydrocarbon production mix. For the first time in recent decades, oil production
decline in the United States has been reversed and natural gas has become abundant. This has
been mainly due to technological advances in both horizontal well drilling and completion
practices. Despite the technological advances, the understanding of physical mechanisms
governing fluid production from these unconventional resource rocks has been limited. The
available analytical and numerical tools were designed to answer fluid flow problems in
conventional rocks, where basic rock properties and fluid flow characteristics are measured, well
studied and understood. There is still an immense need in the industry to characterize and model
flow in the nanodarcy source rocks better.
3

In recent years, new laboratory measurement techniques have been developed and more
detailed rock characterization studies were conducted to better understand fluid flow in
unconventional tight-gas and oil reservoirs. These measurements and studies indicated the
existence of discontinuous microcracks in matrix, which are thought to be a result of the volume
expansion in the process of hydrocarbon generation from kerogen, in addition to an extensively
connected macrofracture network, which is usually associated with the subsequent tectonic events:
uplifting, faulting, folding, etc. As in conventional naturally fractured reservoirs, macrofractures
are connected over the entire producing interval and possess relatively high conductivity. The
microfractures, on the other hand, are the small discontinuous fractures, which are usually
ignored and not characterized in conventional systems because of their small aperture and poor
permeability (usually due to cementation). For a classical dual-porosity, fractured reservoir, the
effective system permeability is a function of the conductivity and density of the fractures. In
unconventional tight formations, the contribution of the connected macrofractures is similar to the
conventional systems; however, microfractures potentially contribute to the effective permeability
of the matrix depending on their length, conductivity, orientation, and connectedness. In general,
for a given density and orientation, the contribution of microfractures to matrix flow capacity
depends on the relative significance of microfracture conductivity with respect to matrix
permeability (Hyman et al. 1991, Slatt et al. 2011) and whether microfractures communicate with
the macrofracture network. Considering the fact that microfractures are mostly disconnected, the
latter implies that only the microfractures closer to the surface of the matrix will contribute to the
flow capacity. Matrix microfractures are usually partially cemented and discontinuous and lack
the conductivity to be important for fluid flow in relatively high-permeability matrix. For
unconventional, tight reservoirs with nanodarcy matrix permeability (such as Eagle Ford, Bakken,
Barnett, Marcellus, etc.), microfractures have the potential to play a role in production. In this
research, a model is developed and used to investigate the effect of matrix-surface stimulation
caused by the existence of surface-connected microfractures and its contribution to fluid transfer
from the matrix medium to the fracture network.
Additionally, prevailing models of flow in unconventional shale reservoirs, whether they
consider fluid flow in both fractures and matrix or incorporate the effect of fluid transfer from
matrix to fractures with dual-porosity idealization, assume that the main contributor of the fluid
transfer is the Darcy flow in the matrix induced by the pressure differential between the matrix
and fracture. Fundamental considerations of Darcy flow, however, reveal that fluid movement in
the nanodarcy shale matrix should be negligible for practical periods of time, unless exaggerated
4

or unrealistic importance is assigned to other parameters to match the sustained production levels
observed in practice. Therefore, a more detailed look at the flow in matrix is essential for better
understanding of the contribution of shale matrix to production.
Another aspect of the fluid flow in unconventional rocks is stress-dependent permeability
reduction, especially in the microfractures and macrofractures. Field tests suggest that these tight
formations would not produce substantial gas or oil, even when developed with horizontal well
technology, without effective hydraulic fracturing techniques. Production, in part, has become
economically feasible with the extensive injection of stimulation fluid/sand to create and keep
these fractures of different scales propped open during the production. It is reasonable to expect
that the micro and macrofractures would lose their conductivity, even though the macrofractures
might be effectively propped, through embedment. To account for the closure of natural fractures
under pressure drop, stress-dependent permeability in natural fractures should be considered. In
this research, an improved dual-mechanism dual-porosity formulation is developed that takes into
account the transient Darcy and diffusive flows in the shale matrix and accounts for the closure of
natural fractures under pressure drop.
1.3 Hypotheses
Hypothesis 1; Based on the existing knowledge of the characteristics of unconventional
reservoir rocks and preliminary modeling results, it is hypothesized that unconnected
microfractures at the matrix surface cause an additional stimulation effect and significantly
improve flow potential in extremely tight rocks. It must be possible to incorporate the impact of
this stimulated zone in dual-porosity models of production from shale reservoirs with fractured
horizontal wells.
Hypothesis 2; Darcy flow in shale matrix cannot solely account for sustained production
from shale reservoirs. Due to the existence of a large volume of nanopores in shale matrix, it is
hypothesized that Knudsen diffusion (slip flow) has a significant contribution on long-term shale-
gas production performance. On the other hand, stress-dependent permeability reduction in
natural fractures is expected to take a toll on the long-term productivity. It must be possible to
incorporate the impact of Knudsen diffusion and stress-dependent permeability reduction in
natural fractures in dual-porosity models of production from shale reservoirs with fractured
horizontal wells.
Hypothesis 3; Coupling flows in nanodarcy matrix and highly conductive fractures
requires a definition of interface boundary conditions which are different from those used in
5

conventional fractured reservoir modeling. Considering improved physics of flow and taking into
account the higher order of flow equations in nanopores, it should be possible to improve the
physical definition of matrix-to-fracture fluid transfer in unconventional reservoirs.
1.4 Objectives
The purpose of this research is to investigate the impacts of unconnected microfractures
on matrix surface, Knudsen diffusion (slip flow) in nanopores, and stress-dependent permeability
reduction in natural fractures on the overall productivity of unconventional reservoirs. The most
important expectation from this research is an improved description of fluid transfer from matrix
to microfractures and then into macrofractures though the various mechanisms indicated above.
The objectives of the research are stated below:
1. Review the existing dual-porosity models from the perspective of shale-reservoir
applications.
2. Review the correlations for stress-dependent natural-fracture permeability.
3. Develop an analytical model and computational code that incorporate unconnected
microfractures on matrix surface in a dual-porosity model.
4. Develop an analytical model and computational code that incorporate Knudsen diffusion
in matrix nanopores and stress-dependent permeability reduction in natural fractures in a
dual-porosity model.
5. Compare the results of these models to those of conventional dual-porosity models to
delineate the effects of microfractures, Knudsen diffusion, and stress-dependent
permeability reduction in fractures.
6. Incorporate the three factors mentioned above into transfer functions used in dual-
porosity models.
1.5 Phases of the Research
This research has consisted of exploratory, constructive, and verification phases:
Exploratory Phase: In the exploratory phase, the problem was defined, including
(i) the statement of the hypotheses,
(ii) investigation of existing knowledge and data,
(iii) definition of the specific objectives and the boundaries of the research, and
(iv) selection of the research methods.
6

Constructive Phase: During the constructive phase,
(i) an analytical model was developed to study and understand the effect of microfractures
on matrix surface on the overall productivity of unconventional reservoirs,
(ii) an analytical model was developed to study and understand the effect of Knudsen
diffusion (slip flow) and stress-dependent permeability reduction in natural fractures on
the overall productivity of shale-gas reservoirs.
Verification Phase: The final phase of the research comprised:,
(i) the verification of the analytical models developed in the research, and
(ii) interpretation of the results.

1.6 Method of Study
Analytical methods have been used in the research. Two analytical models have been
developed in the constructive phase to model flow toward fractured horizontal wells in
unconventional, ultratight reservoirs.
The first model incorporates unconnected microfractures at the matrix surface. To
construct the analytical model, solutions to the diffusion equation were derived for the matrix,
microfracture, and macrofracture media and then coupled. The analytical model was derived in
the Laplace transform domain. Furthermore, to linearize the gas diffusion equation, the
pseudopressure approach was implemented. The computational code was written in Fortran 90.
The second analytical model incorporates Knudsen diffusion (slip flow) and stress-
dependent fracture permeability reduction in addition to Darcy flow in the shale matrix. Similar
to the first model, this model also consists of a coupled matrix-fracture flow solution in the
Laplace transform domain. To obtain the solution, pseudopressure linearization was applied to the
diffusion equation and pressure-dependent fracture permeability was also incorporated. The
computational code was written in Fortran 90.
Field and simulated data have been used in the verification of the analytical model.
Results have been analyzed and interpreted to determine the effects of unconnected
microfractures at the matrix surface, Knudsen diffusion, and stress-dependent permeability
reduction in natural fractures on the overall productivity of unconventional reservoirs.
7

CHAPTER 2
LITERATURE REVIEW
Solutions of fluid flow problems can be obtained by numerical methods or analytical
methods under favorable conditions. Numerical methods, although mathematically approximate,
are more powerful and appropriate for multiphase flow and complex reservoir architecture.
Analytical methods, on the other hand, are rigorous but only applicable for single phase flow and
certain forms of heterogeneity, such as natural fractures and layering. Currently, most
unconventional gas and oil reservoirs in source rocks are modeled with emphasis on smaller-scale
heterogeneities without description of larger-scale geologic variations. To be specific, the most
important heterogeneity in these systems is the existence of stimulated rock volume (SRV), which
consists of a (natural) fracture network. Although it is possible to use numerical models with
discrete fracture network models to represent the SRV, the detailed data to describe fractures in
SRV are nonexistent and impractical to collect, except for qualitative inferences from
microseismic data. An alternative is to use the dual-porosity idealization, which is amenable to
analytical modeling. This is the approach that is taken in this research. Analytical solutions
presented in this research build upon the previous work on fluid flow in naturally fractured
reservoirs and improve them by incorporating the impact of unconnected microfractures at the
matrix surface, Knudsen diffusion, and stress-dependent permeability reduction in natural
fractures.
Analytical solutions of the diffusion equation have been obtained for a variety of
reservoir and well configurations in the literature. Van Everdingen and Hurst (1949) presented
mathematical derivation of fluid flow problems in porous media and introduced Laplace
transformation to petroleum literature. Gringarten and Ramey (1973) introduced Green functions
and sources and sinks for the solution of unsteady flow in porous media. An even a wider range
of solutions had been presented earlier by Carslaw and Jaeger (1959) in the context of heat
conduction in solids. These solutions have been extended to applications in the Laplace
transformation domain by Ozkan and Raghavan (1991). The use of numerical Laplace inversion
algorithms (like Stehfests algorithm, 1970) enhanced the utilization of the analytical solutions.
One of the requirements for the analytical solutions to be applicable is the linearity of the
flow problem which is violated for gas flow in porous media. This is because gas properties are
not independent of pressure; however, this problem may be alleviated by defining a real-gas
pseudopressure, as suggested by Al-Hussainy and Ramey (1966), and Al-Hussainy et al. (1966).
8

Analytical models are usually applied to homogeneous reservoirs but naturally fractured
reservoirs are one of the few exceptions to this rule. To incorporate naturally fractured reservoirs
into numerical and analytical reservoir flow models, dual-porosity models have been proposed in
the literature. Warren and Root (1963) suggested that fluid transfer from matrix to natural
fractures can be assumed to take place under pseudosteady flow conditions. In this idealized
model, hydrocarbon storage (storativity) is mainly controlled by the matrix and it is minimal in
the fractures. Flow capacity, however, is controlled by the fractures and matrix permeability
impact is minimal. He also introduced a shape factor based on the assumption of isotropic, cubic
shaped matrix blocks (sugar cube) that controls the transmissivity between matrix blocks and
fractures surrounding them. Kazemi (1969) proposed a slab model using a similar solution
method to Warren and Roots model and introduced transient flow conditions between matrix and
fracture systems. de Swaan (1976) also presented a transient model that explicitly accounts for
the matrix and fracture dimensions and flow properties. Kazemi et al. (1976) further improved the
solutions for naturally fractured reservoirs by adding the multiphase flow concept and
numerically solved the problem in three dimensions. He also presented a new shape factor in his
solutions.
Triple porosity models for radial systems were first introduced by Liu (1981, 1983) and
Abdassah and Ershaghi (1986) where there are two matrix systems with different properties
feeding a single fracture under unsteady-state flow conditions. Later Al-Ghambi and Ershaghi
(1996) introduced a similar system with two fracture systems with different properties and a
single matrix system. In their first model, one of the matrix layers defined by Abdassah and
Ershaghi (1986) was replaced by a microfracture layer and had different properties than the
macrofracture layer. In the second model they defined a matrix system that only feeds the
microfracture system and the microfracture system feeds the macrofracture system. They
assumed pseudosteady-state flow conditions for their solutions. Both fracture systems were
connected to the wellbore. Dreier (2004) extended these solutions by assuming transient flow
conditions between the microfractures and macrofractures while the flow between matrix and
microfractures is still pseudosteady state.
El-Banbi (1998) introduced linear dual-porosity solutions in fractured reservoirs and
Bello (2009) applied these solutions to horizontal wells in unconventional fractured formations.
Bello and Wattenbarger (2008, 2009, 2010) used these linear-flow models to analyze horizontal
shale-gas wells and identified five flow regions based on constant-pressure production solution.
Ozkan et al. (2009) and Brown et al. (2009) introduced a trilinear flow model for a horizontal
9

well with multiple finite-conductivity hydraulic fractures in a stimulated reservoir volume (SRV).
The SRV in the trilinear model consisted of macro fractures and matrix (modeled as a dual-
porosity medium) between the wellbore and the non-fractured outer reservoir. The hydraulic
fractures, inner reservoir between them (SRV), and the outer (non-fractured) reservoir were
modeled as linear-flow regions and the contiguous system of the three regions gave rise to a
trilinear-flow model. Al-Ahmadi (2011) extended the triple-porosity models (matrix,
microfractures and macrofractures) introduced by Bello and Wattenbarger (2008, 2009, 2010) by
considering pseudosteady state and transient flow conditions between the matrix and fracture in
different media for different flow geometries (radial and linear). In this research, a transient dual-
porosity model is used. Justification for this choice is based on the tight permeability of the shale
matrix, which requires impractically long flow periods to reach pseudosteady flow conditions in
the matrix (Brown et al. 2009).
The early models inherently assumed rock properties did not change with pressure
depletion, although such changes could lead to erroneous outcomes if not accounted for properly.
Pedrosa (1986) constructed type curves for oil and gas flow investigating the pressure-transient
response in stress sensitive formations. He introduced the concept of fractional permeability
change with unit change in pressure, which is equivalent to exponential variation of permeability
with pressure. This was a simplified model that expanded the diffusivity and the permeability
modulus in a Taylor series which limited its applicability. Celis et al. (1994) presented an
analytical model that was constructed to handle the unsteady and pseudosteady state flow of oil
reservoirs and considers flow in naturally fractured, stress-sensitive reservoirs.
Best and Katsube (1995) discussed shales as seals to the hydrocarbon accumulation due
to very tight permeability and the depositional structure of the mineral grains. This tight nature of
the permeability also creates an environment for overpressure. Permeability that could create
overpressure is considered to be in the nanodarcy range. Additionally, tectonic stresses,
temperature increase with depth, and released water from clays in a closed system cause
disequilibrium compaction due to trapped water in pores and create conditions for overpressure.
Kerogen contained in shales is also considered as a potential reservoir. Guegen et al. (2009)
documented the conditions required to create vertical and horizontal microcracks as a function of
mean stress and differential stress. Additionally they made a case for crack-like pores in shales
where the rocks are fine-grained sedimentary and anisotropic.
10

Min et al. (2004) studied the two main mechanisms that cause fracture
aperture/transmissivity. These mechanisms were normal stress induced opening/closure and
shear induced dilation. When the horizontal to vertical stress ratio is low, the impact of shear
dilation appears to be low. In additional experimental studies, Gutierrez et al. (2000) observed
that in de-mineralized fractures in shales, fractures are never closed completely under different
effective normal stresses and shear displacement. These stresses, however, reduce the fractured
permeability significantly.
Raghavan and Chin (2000) coupled fluid flow and geomechanics and presented a model
that considers stress-dependent permeability. In their analysis they considered pressure transient
problems in stress-sensitive reservoirs with nonlinear elastic and plastic constitutive behavior.
Additionally, in 2004, Raghavan and Chin proposed correlations for a variety of rock types that
could be used in evaluating productivity changes in stress-sensitive reservoirs.

11

CHAPTER 3
MICROFRACTURES, DIFFUSIVE FLOW, AND STRESS-DEPENDENT NATURAL
FRACTURE PERMEABILITY
This chapter presents a summary of the existing knowledge and interpretation of the
microfracture generation mechanisms in source rocks, Knudsen diffusion, and stress-dependent
fracture permeability reduction.
3.1 Microfractures in Source Rocks
Substantial volume changes occur in resource rocks during hydrocarbon generation. As
shown in Figs. 3.1. and 3.2., taken from Meissner (2007), with burial depth increase (i.e., pressure
and temperature increase) and over time, original immature kerogen reaches the maturity level
and starts cracking into oil and/or gas (depending on the kerogen type). During this conversion
process, the volume of kerogen spent is substantially smaller than the hydrocarbon volume
generated. This volume imbalance creates excessive pressures, even enough to lift the
overburden pressure. The quality of the seal would then determine the complexity of pathways
(microcracks or microfractures) that will be created during the hydrocarbon migration process
and the overpressure remaining in the system.
Another important mechanism that creates over or underpressure is the impact of the
temperature on the fluids that are in the system, including water (Fig. 3.4. Hunt, 1990). Fig. 3.5
shows theoretical change in water pressure caused by temperature changes with volume
remaining constant (Meissner, 2007). Only a 13C increase in temperature would create 1439 psi
pressure increase in the constant water volume. Similarly, a cooling of the system by 13C would
create a 1224 psi vacuum. This indicates the significant amount of pressure change that a source
rock system would go through during the hydrocarbon generation. Depending on the strength of
the seal rock, these pressure changes could lead to microcracks or microfractures.
The pressure and temperature changes that the source rocks are exposed to over
geological time are due to the burial and uplifting of these formations. These events expose the
rocks to additional stresses other than the lithostatic stress. These are tensional (stretching),
compressional (squeezing), and shearing (side to side shearing) stresses. The rock fabric (ductile
or brittle) would then dictate how the rock would break and how complex of a microfracture
network would form.
12


Figure 3.1 Volume changes taking place in kerogen and generated hydrocarbons during
increasing thermal maturity (Meissner, 2007).

Figure 3.2 Volume changes taking place in oil being cracked to more thermally stable
hydrocarbon species plus residual solid dead oil or coke during increasing thermal maturity.
(Meissner, 2007).
13


Figure 3.3 (A) Pressure/ depth gradient from normal pressure to overpressure across hydraulic
seal surrounding overpressured rock section. (B) Pressure/ depth gradient for similar
underpressured rock system (Hunt, 1990).

Figure 3.4 Theoretical change in water pressure caused by temperature changeswith volume
remaining constant. The Aquithermal effect (Meissner, 2007).

14

At present day stress conditions, most or all of these microfractures could be healed or
closed but exist in the reservoir as points of weakness. Engelder at al. (2008) documented open
microfractures propagating between kerogen particles (Fig. 3.5). Soeder et al. (2009) also
documented partly mineralized fractures in Marcellus Shale in Pennsylvania (Fig. 3.6). Even the
open microfractures at present day would not play a significant role in hydrocarbon flow since
they are not as effectively connected over long distances as the macrofractures. In the absence of
these macrofractures, the horizontal wells drilled in these resource rocks do not produce any
commercial amount of fluid without effective stimulation. Microseismic studies suggest that only
after multi-stage hydraulic fracturing, a complex network of fractures is formed.
Fluid flow from these ultratight reservoirs is by virtue of a stimulated reservoir volume
(SRV) around the well, which consists of a network of natural or induced fractures forming a
continuum within the matrix. This system may be idealized by an appropriate dual-porosity
model as shown in Fig. 3.7. As explained above, ultratight matrix possesses microfractures,
which could be a result of the hydrocarbon generation in source rocks (expulsion cracks), past
tectonic events (folding and faulting), burial and uplift, and drilling and hydraulic fracturing. For
modeling purposes, however, more important are the shorter lengths and the disconnectedness of
these microfractures. The effect of an isolated microfracture embedded in the matrix is negligible
as it only causes minor and local disturbance to flow lines. Microfractures reaching the surface of
the shale matrix (that is, the microfractures intersected by the macrofracture network) may
enhance the local flow capacity and fluid transfer from matrix to macrofracture network
considerably. With this interpretation, the effect of the microfractures can be incorporated by a
fractured surface-layer of the shale matrix as shown in Fig. 3.8. The thickness of the surface layer
depends on the lengths of the surface-connected microfractures.
3.2 Knudsen Diffusion
Flow mechanisms that govern flow in resource rocks have become of great interest
recently with increased commercial production. One of the major factors determining the
productivity of shale reservoirs is the existence of a natural fracture network. In most cases, a
question arises about the contribution of the shale matrix. Unfortunately, a complete
understanding of fluid transfer from shale matrix to fracture network has not yet been achieved.
Current studies, whether they model fluid flow in both fractures and matrix or incorporate the
effect of fluid transfer from matrix to fractures with dual-porosity idealization, assume that the
main contributor of the fluid transfer is the Darcy flow in the matrix induced by the pressure
differential between the matrix and fracture. Fundamental considerations of Darcy flow, however,
15



Figure 3.5 (A) and (B): microfractures propagating between kerogen particles (Engelder et al.
2008).
16


Figure 3.6 Partly mineralized fracture from Marcellus Shale in Pennsylvania (Soeder et al. 2009).


Figure 3.7 Dual-porosity idealizations of naturally fractured reservoirs (Najurieta, 1980).
17


Figure 3.8 Composite matrix model to idealize flow in microfractured shale matrix.
reveal that fluid movement in the nanodarcy shale matrix should be negligible for practical
periods of time, unless exaggerated or unrealistic importance is assigned to other parameters to
justify the sustained production levels observed in practice. Therefore, a more detailed look at the
flow in the matrix is essential for better understanding of the contribution of shale matrix to
production.
Compared to conventional rocks, source rocks are extremely tight, with permeabilities in
the low micro-to nanodarcy range. These rocks, until a few years ago, were considered seal rocks
where no commercial production would be possible (today, without proper drilling and
stimulation techniques and without the existence of extensive connected natural fractures,
commercial production is still not viable). These rocks have such small pore throats that
buoyancy pressure ("!gh) of the two fluids (water and the hydrocarbon fluids) that exist today
would not be enough for the passage of these fluids through them (Nelson, 2009).
As shown in Fig. 3.9 (Nelson, 2009), measurements reveal that pore throat sizes in source
rocks are in the 0.005 to 0.1 "m range, while for tight sandstones they are in the range of 0.03

to
2 "m and for sandstones they are greater than 2 "m. Hydrocarbon molecules, such as paraffines,
asphaltenes, ring structures, and methane, however, have particle sizes between 3.8 (0.00038
"m) to 100 (0.01 "m), asphaltenes and methane being the largest and the smallest, respectively.
It is evident in Fig. 3.7 that gas particles, oil particles (unless very complex), and water can fit
through the pore throat sizes that dominate the source rocks. For a sense of the porosity-
18

conductivity relationship, it should be noted that an order of magnitude change in pore-throat size
corresponds to two orders of magnitude change in permeability.

Figure 3.9 Composite Sizes of molecules and pore throats in siliciclastic rocks on a logarithmic
scale covering seven orders of magnitude (Modified from Nelson, 2009, by Terri Olson, 2012).
Nelson (2009) prepared Fig. 3.9 by using the data from Jurassic Cretaceous Shelf Shales
(8-17 nm pore-throat diameter), Devonian Appalachian Basin Shales (7-24 nm pore throat
diameter), Pliocene Beaufort-Mackenzie Basin Shales (9-24 nm pore-throat diameter) and
Pennsylvanian Anadarko Basin Shales (20-160 nm pore-throat diameter). In Fig. 3.10, pore-
throat radius distribution is given for a Barnett-Shale well. Eighty percent of the pore throats
have a radius of less than 0.005 "m. This particular sample was taken at 8,094 ft in Mildred
Atlas # 1 well of Enre Corporation (Chevron) in Johnson County, (modified from Bowker, 2007b
by Bruner et al., 2011). This is in line with the pore-throat diameters reported by Nelson (2009).
Fig. 3.11 presents pore-throat histogram data from another Barnett Shale well (Blakely #1)
calculated from mercury capillary measurement data. It can be seen that calculated pore-throat
diameters are in the 5-15 nanometer range (Loucks et al. 2009). Porosities and permeabilities
#$%$&
&'()*+,
-./*+,

19

shown in Fig. 3.11 are from core-plug measurements. Loucks presented at the same study some
scanning electron microscope (SEM) images from the same well (Fig. 3.12) that support the pore-
throat diameters shown in Fig. 3.11. In practical terms, macroporosity corresponds to less than
0.05% of the total porosity in these unconventional nanopore systems.
Recent technological enhancements in pore imaging and pore level property
measurements present a big challenge in terms of quantifying what each measurement technique
provides. In plug level and crushed sample protocol (CSP) measurements, pore properties
(porosity, saturation, and throat sizes) are measured at vastly different scales. In imaging
techniques, pore-body sizes are measured by using different approaches (Fig. 3.9) and the
resolution of the imaging techniques and the numerical techniques used to convert the images to
quantities dictate the outcome of these measurements. Additional assumptions are made to
construct 3-D flow models to calculate permeabilities from these images.
Fig. 3.13 demonstrates how variable measurement techniques conducted by two
laboratories measure porosity for core data taken from Niobrara formation. Lab A measured
porosity by using crushed sample protocol (CSP) and by conventional methods (at plug level)
after Dean Stark extraction. As can be seen in the plot, porosity measurements taken from the
same depth vary between two to five times. Later, cores taken from same depths were tested by
Lab B and they measured porosity at plug level as in the state that they received the core plugs.
As evident in Fig. 13, Lab Bs measurements consistently suggest lower porosity compared to
Lab A. After that, Lab B extracted (cleaned) the cores for 4 weeks with toluene and re-measured
the porosities. Measured porosities significantly improved. Later, they continued the cleaning
process for another four weeks with chloroform-methanol azeotrope. The measured porosities
further increased. As the last measurement technique, Lab B applied CSP and cleaned the
crushed plug particles for five days by toluene and five days with azeotrope and then measured
porosities. After this extensive cleaning process, both plug level and CSP measurements yielded
similar results. This process demonstrates that cleaning processes and lab techniques result in
significantly different porosity estimates.
The continuum hypothesis is an approximation that assumes that fluids are continuous,
while in reality they are formed by molecules that collide with each other and the surface of the
solids around them. It is assumed that properties like pressure, temperature and velocity of the
fluid within a control element volume vary continuously and can be averaged, while the fluid
itself is made up of discrete molecules (http://en.wikipedia.org/wiki/Fluid_mechanics). Flow in
20


Figure 3.10 Results of mercury-porosimetery analysis of the Barnett Shale (Modified from
Bowker, 2007b by Bruner et al., 2011).

Figure 3.11 Pore-throat diameters calculated from four capillary-pressure sample analyses of
Barnett Shale in the Blakey #1 well (Loucks et al. 2009).
21

pipelines is governed by the Navier-Stokes equation when the continuum hypothesis is valid.
When the diameter of the flow conduits in porous media becomes comparable to the mean-free-
path of the fluid molecules, then the continuum hypothesis is not applicable anymore (Beskok,
1996). Knudsen number may be used to determine which flow regime would be applicable when
the continuum hypothesis is no longer valid (http://en.wikipedia.org/wiki/Fluid_mechanics).
Knudsen number is defined as the ratio of the distance a particle would move (the mean free path
of the fluid molecules) before it hits another particle of its own size diameter and the
representative dimension of the flow conduit (diameter of a pore channel) (Roy et. al. 2003). It
determines how much rarefaction effect occurs when gases flow through micro channels.
As shown in Fig. 3.14 (Roy et. al. 2003) when Knudsen number is smaller than 0.001,
continuum flow occurs and between 0.1 and 0.001, slip flow prevails in the system. At or above
unity for the Knudsen number, free molecular flow exists and statistical methods are appropriate
to describe flow. As Knudsen number increases, rarefaction effects become so dominant that the
continuum hypothesis is not applicable anymore.
Darcys law is commonly used in the oil industry to model fluid flow in porous media.
This law includes the continuum hypothesis and states that apparent fluid velocity is directly
proportional to the pressure differential applied. Similar to the Navier-Stokes equation, one of the
assumptions of this law is that fluid velocity at the pore walls (pipe or channel surfaces for the
Navier-Stokes equation) is zero (no-slip flow). Klinkenberg in 1941 identified that at low mean
flowing pressures, the diameter of the gas-flow path approaches the length of the mean free path
of gas molecules, which would impact air permeability measurements. As defined earlier by
Knudsen number, this would cause slip flow at pore walls. As it is shown in Fig. 3.15, when gas
and liquid flow at high mean flowing pressures, flow is laminar, the Darcys law is valid, and
there is no-slip flow at the walls. At low flowing pressures, however, Darcys law is not valid
anymore and slip flow exists where the flow velocity is nonzero at the walls. Figure 3.16
presents the velocity spectrum changing from very high to very low. At very high velocities, due
to non-linear flow, Forcheimers equation is used in conjunction with Darcys law. At moderate
velocities Darcys law is used since the flow is laminar. At low velocities, as explained earlier,
slip flow causes non-linear flow.
It must be noted that the Klinkenberg effect used to correct lab measurements considers
only one type of slip flow and constitutes an approximation. In a more general sense, Knudsen
diffusion can be used to formulate slip flow in porous media. In this research, Ficks law of
22


Figure 3.12 Secondary Electron Images showing variations of morphology in organic matter
(Loucks et al. 2009).
(a) Very small (18-46 nm diameters), nearly spherical nanopores, porosity 5.2%, depth 2186.4 m.
(b) Larger nanopores (550 nm diameter) pillar like nanopores, depth 2167.4 m.
(c) Tube-like pore throats connecting elliptical pores (<20nm), depth 2167.4 m.
(d) More tube-like pore throats connecting elliptical pores (<20nm), depth 2167.4 m.


Figure 3.13 Extended Extraction Measurements from a Niobrara Core.
23

diffusion is used to account for Knudsen diffusion in porous media (Ertekin et. al. 1986). Ficks
law defines molecular flow as transport of matter from one point to another in a system through
random molecular motions and it states that mass flow rate is directly proportional to
concentration gradient. Ertekin et. al. modeled slip flow in nanopores by using Ficks law and
introduced a slip velocity as a function of the pressure gradient. Then, they defined a total
(effective) velocity as the linear sum of the Darcian and slip velocities. To highlight the similarity,
they re-casted the effective velocity in Klinkenbergs form. This approach yielded an effective
permeability which included a correction term for the Darcy velocity. The new correction term
obtained by Ertekin et al. had the same form of the Klinkenberg correction but, unlike
Klinkenbergs constant correction term, it was pressure dependent. They highlighted the
differences between the Klinkenbergs constant slippage and their dynamic slippage on practical
examples (Ertekin et al. 1986).
In recent studies (Javadpour et al., 2007, Javadpour, 2009), gas flow in shale matrix has
been described by Knudsen diffusion and slip flow in the nanopores, Darcy flow in the
micropores, desorption from the surface of kerogen, and diffusion in solid kerogen (Fig. 3.17).
Fig. 3.18 summarizes the flow regimes, pore diameter ranges and permeability ranges for
conventional, tight and shale oil and gas systems.
One of the objectives of this research is to incorporate this more detailed description of
flow in shale matrix to modeling of production from fractured shale-gas reservoirs. Desorption in
shale reservoirs has been likened to that in coalbeds where gas desorption takes place from the
surface of the coal matrix to the cleat system. The major difference in shale-gas reservoirs is that
desorption takes place from the surface of the organic content (kerogen) embedded in the shale
matrix to nanopores. Therefore, to describe desorption in shale matrix, in addition to the standard
parameters, such as the volume and maturity of the organic content and the Langmuir isotherms,
the distribution of kerogen, pressure profile, and the exposed surface area of nanopores in the
shale matrix should be considered. This information is currently nonexistent or incomplete and,
thus, incorporating desorption into gas-flow model for shale matrix is deferred until later. Hence,
this research focuses on Darcy and diffusive flow processes.
3.3 Stress-Dependent Fracture Permeability Reduction
Field tests suggest that source rocks would not produce substantial gas or oil, even when
developed with horizontal well technology, without effective hydraulic fracturing techniques.
Fractures could exist at different scales, micro or macro, and could be man-made or natural. All
24


Figure 3.14 Knudsen number regimes and governing equations (Roy et. al. 2003).


Figure 3.15 (A) High mean flowing pressure (B) Low mean flowing pressure


Figure 3.16 Flow regimes, pore diameter ranges and permeability ranges for conventional, tight
and shale oil and gas systems.
25


Figure 3.17 Gas flow mechanisms at different scales (Javadpour et al., 2007).



Figure 3.18 Flow regimes, pore diameter ranges and permeability ranges for conventional, tight
and shale oil and gas systems (Figure from Javadpour et al. 2007).
26

of them play a vital role in hydrocarbon production from these very tight formations. Production,
in part, became economically feasible with the extensive injection of stimulation fluid/sand to
create and keep these fractures of different scales propped open during the production. It is
reasonable to expect that these micro and macrofractures would lose their conductivity, even
though they might be effectively propped open, through fracture closure or embedment. To
account for the closure of natural fractures under pressure drop, stress-dependent permeability in
natural fractures should be considered.
Important characteristics of fractures are aperture, length, permeability and porosity.
Fracture orientation and spacing (density) also play a major role in horizontal well placement and
productivity. Early to intermediate flow periods are dominated by the fractures where better
conductivity exists. This causes a rapid pore pressure drop, hence a substantial increase in net
effective stress. This leads to closure of the fracture aperture causing permeability reduction.
Additionally, temperature and strain could cause changes in the mechanical properties of the
porous and fractured media. The two main mechanisms that cause fracture
aperture/transmissivity change are normal stress-induced closure and opening- and shear-induced
dilation. When the horizontal to vertical stress ratio is low, the impact of shear dilation appears to
be low (Min et al., 2004). This has been an important study area in rock mechanics.
Raghavan and Chin (2004) proposed correlations for stress sensitive reservoirs that
predict permeability reduction based on experimental data. They have proposed three correlations
that are in exponential, linear and power law form. In this research, exponential pressure
dependence formulation is used to model the stress-dependent permeability reduction in fractures.


27

CHAPTER 4
ANALYTICAL FLOW MODELS
This chapter introduces the derivation of the analytical models. Two analytical models
have been developed to model flow toward fractured horizontal wells in shale reservoirs.
The first model incorporates unconnected microfractures at the matrix surface. To
construct the analytical model, solutions to the diffusion equation are derived for the matrix,
microfracture, and macrofracture media and then coupled. The analytical model is derived
mathematically in the Laplace transform domain. Furthermore, to linearize the gas diffusion
equation, the pseudopressure approach is implemented.
The second model incorporates Knudsen diffusion in matrix and stress-dependent
permeability in macrofractures (this model does not consider matrix microfractures). To construct
the analytical model, solutions to the diffusion equation are derived for the matrix and fracture
media and then coupled. The analytical model is derived mathematically in the Laplace transform
domain. Furthermore, to linearize the gas diffusion equation and incorporate the pressure-
dependent fracture permeability, the pseudopressure approach is implemented.
In Section 4.1, general assumptions that are valid for both models are discussed and the
details of both models are introduced in the subsequent sections.
4.1 General Assumptions, Definitions of Dimensionless Variables and Model Parameters
For derivation of analytical equations several assumptions are made. It is assumed that
fluid is single phase and flow is isothermal. The permeability tensor is assumed to be isotropic.
The reservoir has both homogenous and naturally fractured segments.
Because the intent is to derive an analytical model, the linear form of the diffusivity
equation is used (i.e., the liquid-equivalent flow of a fluid is assumed). For analysis of gas wells,
the usual approach is followed and the liquid-flow analogy of Al-Hussainy et al. (1966) is
incorporated through the pseudopressure transformation:
( ) ( ) ( ) 2
i
p
i i
p
g
p
m p m p m p dp
z
!

"
" # = $ =
%
, (4.1.1)
and ! is defined depending on the context in the following sections. This pseudopressure
definition linearizes the diffusion equation for the flow of a real gas and pressure-dependent
permeability.
28

In the following definitions, the subscripts , , , , f m mc mm mf and ms indicate the property
of macrofracture, matrix, matrix core, surface-layer matrix, matrix microfracture, and matrix
surface layer, respectively. Also the subscript F is used to indicate that the property belongs to
the hydraulic fractures, O to represent the outer-reservoir (beyond Stimulated Reservoir Volume,
SRV) properties, and D to denote dimensionless variables. The subscripts , , , , , g i sc e wf and t
stand, respectively, for gas, the initial condition, standard (surface) conditions, exterior boundary,
flowing wellbore, and total-system property. In the definitions of dimensionless variables, an
arbitrary characteristic length L is used as the scaling quantity ( Lis later chosen to be the
hydraulic fracture half length,
F
x ). The bar sign is used over the notation to indicate the Laplace
transformation of the property and the following dimensionless variables are defined,
5
1.988 10
, where , , , , or
fi ft sc
D
sc sc
k h T
m m f m mc mf ms
q p T
! !
!
"
#
= $ = , (4.1.2)
( ) ( )
( ) , ,
, ,
mD D D D f msc sc
r R t
q r R t n q q =
, (4.1.3)
where
msc
q is the volumetric flow rate, at standard conditions, crossing a radius r in the matrix
block and
f
n is the number of natural fractures in the pay thickness.
( )
4
2
3
is a characteristic length
, where 2.637 10 for in hr
6.328 10 for in D
c fi
D
c
f tf g
i
L
k
t t t
c L
t
!
!
"
#
#
$
%
$ = &
' %
=
'
%
& %
( (
, (4.1.4)
D
r
r
L
= , (4.1.5)
and
D
R
R
L
= . (4.1.6)
In Sections 4.2 and 4.3, assumptions specific to the individual flow models developed in
this work will be discussed in more detail.
4.2 Dual-Porosity Model with Matrix, Microfracture, and Macrofracture Media
In this section, an analytical model to represent flow in an unconventional, ultratight
formation is derived. The model considers a dual-porosity system consisting of a macrofracture
network and a microfractured matrix. Spherical geometry is considered for the shale matrix
blocks encircled by a microfractured surface layer; the extension of the model to different matrix
geometries is straightforward. The derivations are demonstrated for a radial-flow system with a
29

vertical well as sketched in Fig. 4.1 and the extension is presented to fractured horizontal wells as
they are the most common well completion in ultratight unconventional reservoirs. For fractured
horizontal wells in shale, the tri-linear flow model (Ozkan, 2009, Brown et al., 2009) is used.


Figure 4.1 Schematic of a radial dual-porosity medium with spherical matrix blocks.
Below, a model for flow in the composite matrix (consisting of a microfractured surface
layer and homogeneous core) is derived first. Then, the solution for flow in the macrofracture
network including influx from the matrix is derived and coupled with the matrix flow at the
matrix-fracture interface.
4.2.1 Flow in a Microfractured Spherical Matrix Block
Assume that the spherical matrix blocks are of a composite structure which consists of
two concentric spheres as shown in Fig. 4.2. The interior sphere (the core of the matrix block) is
made of a matrix that may have discontinuous microcracks, which do not significantly contribute
to flow capacity. Therefore, the matrix core acts like a homogenous matrix block. The exterior
sphere (surface-layer of the matrix block) is of the same matrix and has discontinuous
microcracks that are connected to the macrofracture system. Effectively, these microcracks
create an enhanced permeability surface layer, which is the subject of this research. The flow
equations for the core and the surface layer of the matrix block are derived separately and then
the two solutions are coupled at the boundary of the two regions. It is assumed that the
disconnected microcracks in the surface layer create an equally spaced continuous fracture
medium, which could be modeled as a dual-porosity medium where flow capacity is dominated
by the microfractures (this assumption is justified for microfractures in very tight matrix such as
in shale). Therefore, the flow from the matrix core is coupled with the flow in the surface-layer
fractures.
!
"
"
#
$%&!'(
)!%*&+!,
-,../0!,
"12
!1!
#
30


Figure 4.2 Spherical matrix block with a composite structure.
To model flow in the matrix-surface layer, it is assumed that the surface-layer thickness
(
ms m mc
h r r = ! ) is small compared to the matrix-core radius (
mc
r ) and the pressure and flux are
uniformly distributed on the boundaries of the surface layer (
mc
r and
m
r ). Then, one can
approximate the surface layer by a linear system as sketched in Figs. 4.3 and 4.4. Details of the
derivations are given in Appendix A. Here, the final form of the solutions is presented.

Figure 4.3 Representation of the fractured matrix-surface layer by a system of slabs in parallel.

Figure 4.4 Geometry of the matrix and fracture slabs used to represent the matrix surface layer.
The dimensionless pseudo pressure solution for the matrix core in the Laplace-transform
domain is given by
31

( )
( )
( )
( )
sinh
, , , ,
sinh
mcD mDi D
mcD D mD mfD mcD mD
D mDi mcD
r s r
m r R s m r R s
r s r
!
!
= (4.2.1)
where the following is defined
( )
( )
f tf g m
mi i
mDi
fi m tm g f
i
c k
c k
!
"
"
" !
= =

(4.2.2)

The pseudo pressure in this solution is defined by (4.1.1) with 1 ! = and ( ) , ,
mfD mcD mD
m r R s
corresponds to the surface-layer fracture pseudo pressure at the interface of the matrix core and
the surface layer. As shown in Appendix A, the solutions for the surface-layer matrix and the
surface-layer microfractures are given, respectively, by
3
cosh
3
cosh
m
D
m mfDi
msD mfD
m
m mfDi
s
m m
s
!
"
# $
!
# $
% &
' (
' (
' (
) *
=
% &
' (
' (
' (
) *
, (4.2.3)
and
( )
( ) ( )
( )
( ) ( )
( )
( ) ( )
exp 2
,
exp 2
exp ,
m mf m mf m D mcD
mfD D
m mf m mf m mD mcD
m mD D fD mD
u f u f u r r
m r s
u f u f u r r
u r r m R s
! "
# + + #
$ %
=
! "
# + + #
$ %
! "
& #
$ %

(4.2.4)

In obtaining (4.2.4), (4.2.1) and (4.2.3) are used and the following definitions are made:
( )
m m
u sf s = , (4.2.5)
( )
3 1
1 tanh
3
m m mfDi
m
m
mfDi m mfDi
f s s
s
! " #
"
# ! #
$ % & '
( ) *+
= +
( ) *+
) *
( +
, - . /
, (4.2.6)
and
( )
2
2
coth 1
12
mm m
mf mDi mcD mDi mcD
mcD
h
f s r s r
r L
!
" "
# $
= %
& '
, (4.2.7)
where
32

( )
( )
f tf g mf
mfi
i
mfDi
fi mf tmf g f
i
c k
c k
!
"
"
" !
= = . (4.2.8)
Following Serra et al. (1983), it is also defined that
2
2
12
m mm
m
mm mf mf
k h L
h k h
!
" #
$ %
=
$ %
$ %
& '

(4.2.9)

and
( )
( )
t mm
mi
m
t mf
mfi
c h
c h
!
"
!
= . (4.2.10)
The solution for the matrix microfractures given in (4.2.4) is used to derive an expression for the
flux from the matrix blocks into the macrofracture network. The solution for the macrofracture
network is discussed next.
4.2.3 Flow in the Macrofracture Network
A fully penetrating vertical well in a cylindrical reservoir composed of a microfractured
matrix medium and a macrofracture network is considered. To represent the fractured medium,
we use the dual-porosity idealization where flow is dominated by the macrofracture system and
the matrix system supplies fluids into the macrofracture system as pressure drops. It is assumed
that fluid transfer from matrix to macrofractures is through the matrix microfractures only and the
flux from each spherical matrix block is instantaneously and uniformly distributed in one-half the
fracture volume that envelops the spherical matrix block (de Swaan O, 1976). Furthermore, it is
assumed that the fracture volume that envelopes the matrix block has a uniform average thickness
of
f
h , which is small compared with the radius of the matrix block,
m
r .
As shown in Appendix B, under the above listed assumptions, the solution for flow in the
macrofracture network is given, in the Laplace transform domain, by
( )
( )
0
1
D
fD
wD wD
K uR
m
s uR K uR
= , (4.2.11)
where
0
( ) K z and
1
( ) K z are the modified Bessel functions of the second kind of order zero and
one, respectively,
33

( ) u sf s = , (4.2.12)
and
( ) ( ) ( ) 1
m f
f s f s f s ! = " . (4.2.13)
In (4.2.13), the following is defined
( ) ( )
( )
( ) ( )
( )
exp 2
5 exp 2
m mf m mf m mD mcD
mf mD
f
m mm m mf m mf m mD mcD
u f u f u r r
h r
f
u h u f u f u r r
! "
# $
% &
' ' + '
( ( ) *
+ ,
=
- .
+ ,
( ( % &
' + + '
/ 0
) * 1 2
, (4.2.14)
and
2
2
10
mf m
f f m
k r L
k h r
! = . (4.2.15)
The Laplace-domain solution given in (4.2.11) is inverted into the real-time domain by using a
numerical inversion algorithm. For the results presented in this paper, we have used the numerical
Laplace inversion algorithm suggested by Stehfest (1970).
The solution for a fully penetrating vertical well given in (4.2.11) is the same as that for a
vertical well in a conventional dual-porosity reservoir where the matrix is homogenous
(unfractured). However, the definition of ( ) f s given in (4.2.13) is new and reflects the composite
nature of the matrix blocks and the contribution of microfractures on matrix surface. Therefore,
replacing ( ) f s with that given in (4.2.13) incorporates the effect of microfractures on matrix
surface into conventional dual-porosity solutions. The solution for a fractured horizontal well
surrounded by a stimulated (fractured) reservoir volume will be shown later. Next, the solution
will be verified by checking the known asymptotic forms of
( ) f s .
4.2.4 Verification of the Solution
To verify the solution derived above, two asymptotic cases can be checked: the
homogeneous reservoir solution and the conventional dual-porosity solution without matrix
microfractures. The fractured reservoir (dual-porosity) solution derived here should yield the
homogenous reservoir solution when 0
mD mcD
r r = = . The solution given in (4.2.11) is the same as
the homogeneous reservoir solution when u s = , that is, when
( ) 1 f s = . From (4.2.14) and
(4.2.13), respectively,
( )
0
lim
5
mD mcD
mf mD mf
f
r r
mm m
h r f
f s
sh f
= !
" #
$ %
= &
$ %
' (
, (4.2.16)
34

and
( )
0
lim 1
5
mD mcD
mf
mD mf
r r
mm
h
f s r f
sh
!
= "
# $
% &
= +
% &
' (
. (4.2.17)
Also from (4.2.7), one can write
( )
( )
2
0
lim coth 1 0
12
mD mcD
mmD m mD
mD mf mDi mcD mDi mcD
r r
mcD
h r
r f s s r s r
r
!
" "
= #
$ %
= & #
' (
. (4.2.18)
Then, (4.2.17) yields
( ) 1 f s = , which verifies that the solution is the same as that for the
homogeneous reservoir case when 0
mD mcD
r r = = .
The conventional dual-porosity solution with homogenous spherical matrix blocks (no
matrix microfractures) is the same as that in (4.2.11) except for the
( ) f s function given in
Section 4.3 and Appendix C.
( ) 1 1 coth
5
m
mD mD
mf mDi mDi
k s s
f s r r
k s
!
" "
# $ % &
= ' ' ( ) * +
* +
( )
, - . /
. (4.2.19)
It should be shown that the
( ) f s function given in (4.2.13) becomes the same as that in (4.2.19)
when there is no microfractured matrix surface layer; that is, when
mcD mD
r r = . Using (4.2.14),
(4.2.13), and (4.2.7), it can be shown that
( ) lim
5
mcD mD
mf mD mf
f
r r
mm m
h r f
f s
sh f
!
" #
$ %
= &
$ %
' (
, (4.2.20)
( )
0
lim 1
5
mD mcD
mf
mD mf
r r
mm
h
f s r f
sh
!
= "
# $
% &
= +
% &
' (
, (4.2.21)
and
( )
( )
2
2
lim 1 coth
12
mcD mD
mm m
mD mf mDi mcD mDi mcD
r r
h
r f s s r s r
L
!
" "
=
# $
= % %
& '
. (4.2.22)

Substituting (4.2.22) into (4.2.21) yields the same expression for
( ) f s as in (4.2.19). Therefore,
the solution derived here is the same as the conventional dual-porosity solution when
mcD mD
r r = ;
that is, when there are no microfractures in matrix blocks.
In addition to the analytical verification presented above, the solution for the asymptotic
cases was verified without microfractures given in the formulation above and compared with the
35

conventional dual-porosity model without microfractures (Ozkan et al., 2010). The numerical
results for the asymptotic cases are included in the discussion of results under Section 5.1.
4.2.5 Extension to Multiple Fractured Horizontal Wells with Stimulated Reservoir Volume
The trilinear-flow model (Ozkan et al., 2009, and Brown et al., 2009) with the
assumption of spherical matrix geometry (Section 4.3 and Appendix C) will be used here for a
multiple fractured horizontal well in shale. A sketch of the trilinear-flow model is shown in Fig.
4.2.5. The trilinear-flow model considers a set of transverse hydraulic fractures in a stimulated
reservoir volume (SRV), which consists of an extremely tight matrix (such as shale) and a
network of macrofractures. The fractured medium in the SRV is represented by the dual-porosity
idealization. Outside the SRV, the reservoir is made of the same matrix as in the SRV but it is not
fractured.

Figure 4.5 Trilinear-flow model consisting of a fractured inner reservoir (SRV) and a
homogenous outer reservoir (Ozkan et al., 2009, and Brown et al., 2009).
The details of the derivation of the trilinear-flow model are given by Ozkan et al. (2009)
and Brown et al. (2009) for slab-matrix geometry. Appropriate modification of the dual-porosity
parameters for spherical matrix geometry is explained in detail in Section 4.3 and Appendix C.
The dimensionless wellbore pressure for the trilinear-flow model is given by

( )
tanh
c
wD
FD F F
s
m
s
C s
!
" "
= + , (4.2.23)
where
36


2
F
F
FD FD
s
C
!
"
#
= + , (4.2.24)
tanh
2
D
F O O eD
w
y ! " "
# $ % &
= '
( ) * +
, - . /
, (4.2.25)
O
O
RD eD
u
C y
!
" = + , (4.2.26)
( ) tanh 1
O OD OD eD
s s x ! " "
# $
= %
& '
, (4.2.27)
( ) u sf s = , (4.2.28)
where
( ) f s is given in (4.2.19). In (4.2.23),
c
s is the skin factor due to flow choking within the
fracture (Mukherjee and Economides, 1991) given by

ln
2 2
f ft
c
F F w
k h
h
s
k w r
!
" # $ %
= &
' ( ) *
' ( + , - .
. (4.2.29)
To incorporate the effect of microfractures on the surfaces of the matrix blocks into the trilinear-
flow solution given in (4.2.23), the expression for
( ) f s given in (4.2.19) is replaced by the one
given in (4.2.13). Results discussed in Section 5.1 have been obtained by using the trilinear-flow
solution in (4.2.23) with the appropriate expression for
( ) f s in (4.2.13).
4.3 Dual-Porosity Model with Knudsen Diffusion (Slip Flow) in Matrix and Stress-
Dependent Fracture Permeability Reduction
In this section, an improved dual-mechanism dual-porosity formulation that takes into
account the transient Darcy and diffusive flows in the shale matrix is presented. Spherical matrix
blocks are considered in this research, but the formulation can be readily extended to other matrix
block geometries. To account for the closure of natural fractures under pressure drop, stress-
dependent permeability in natural fractures is also considered by using the approach proposed by
Raghavan and Chen (2004). The general formulation presented in this research and the transfer
function for the transfer of fluids from matrix to fracture can be used for numerical or analytical
modeling purposes. The new formulation of matrix to fracture fluid transfer is used with the
analytical trilinear flow model presented by Ozkan et al. (2009) and Brown et al. (2009) for
fractured horizontal wells in shale, and the differences from the conventional formulation are
demonstrated.
37

4.3.1 Introduction of Knudsen Diffusion (Slip Flow)
In this research, a dual-porosity medium with spherical matrix blocks of uniform radius,
m
r , is considered. It is assumed that the pressure at the surface of the matrix blocks is uniform so
that the flow within the matrix is spherical. Using the approach taken by Ertekin et al. (1986), the
radial component of the matrix flow velocity,
rm
v , is defined as the sum of the radial components
of the Darcy velocity,
prm
v , and the slip velocity,
srm
v ; that is,
rm prm srm
v v v = + . (4.3.1)
The radial component of the Darcy velocity in (4.3.1) is given by
m m
prm c
g
k p
v
r
!

" #
$
= % &
% &
' ( $
. (4.3.2)
The radial component of the slip velocity in (4.3.1) is given by the modified form of the Ficks
Law (Ertekin et al., 1986) as follows:
g g
m
srm c
g
M D
C
v
r
!
"
# $
%
= & '
& '
( ) %
. (4.3.3)
In (4.3.3),
g
M is the molecular weight of the gas,
g
D is the diffusivity coefficient for the matrix,
and
m
C is the concentration. Then,
g g
m m m
rm c
g g
M D
k p C
v
r r
!
"
# $
% & % &
' '
( )
= + * + * +
* + * + ( )
, - , - ' '
( )
. /
. (4.3.4)
From the real gas law, molar concentration for single phase flow is
g
m
m
g g
p
C
M zR T
!
= = . (4.3.5)
Thus, (4.3.4) can be written as
( )
g g m m m
rm c
g g g
M D p z k p
v
r R T r
!
"
# $
% & ' (
) ) * *
= + + , - .
/ 0
+ ,
- .
* *
1 2 3 4 ) )
5 6
. (4.3.6)
Because
1 1 g
pc
p p dz p p
r z z p z dp r z r
! "
! "
# # #
$ %
= & = $ %
$ %
$ %
' ( # # #
' (
, (4.3.7)
(4.3.6) can be written as
38

1
g g g
m m m m m
rm c g g c
g g m
c D
k p p k p
v c D
r r k r

! !

" # " #
$ %
$ % $ % $ %
& & &
' ( ' ( ) *
= + = + ) * ) * ) *
) * ) * ) * ' ( ' ( ) *
+ , + , + , & & &
' ( ' ( + ,
- . - .
. (4.3.8)
Following Ertekin et al. (1986), an apparent gas slippage term is defined by
g g g m
am
m
D c p
b
k

= , (4.3.9)
and an apparent matrix permeability by
1
am
am m
m
b
k k
p
! "
# $
= +
# $
% &
. (4.3.10)
The apparent matrix permeability in (4.3.10) is in the form defined by Klinkenberg (1941). As
indicated by (4.3.9), however,
am
b is not a constant; it is a function of pressure. In this research,
the following correlation given by Ertekin et al. (1986) is used to compute the diffusivity constant:
0.67
31.57
g
g
D k
M
= , (4.3.11)
where
g
D is in sq ft/D.
Fig. 4.6 shows the comparison of the apparent matrix permeability,
am
k , and the Darcy
permeability,
m
k , for a 0.8 specific gravity gas of molecular weight 16 lbm/lbm-mol at 120
o
F.
The contribution of diffusive flow is more significant at low pressures and for lower values of
permeability. As shown for
8
10
m
k
!
= md in Fig. 4.6, the difference between the apparent and
Darcy permeabilities may be an order of magnitude. Fig. 4.7 shows the percentage error in
permeability as a result of neglecting diffusive flow in matrix.
4.3.2 Introduction of Stress-Dependent Natural-Fracture Permeability
To account for the closing of fractures when pressure drops in the reservoir, a stress-
dependent permeability is used in this natural-fracture flow model. The main assumption that
governs the flow equations here is that the rock deformation that is considered for the fractures is
perfectly elastic, linear and reversible. Additionally, the stress is distributed uniformly, and the
stresses at the matrix and fracture interface remain constant. Here, following Raghavan and Chin
(2004), the stress-dependent natural-fracture permeability may be defined by
( )
exp
f f
d p
f fi f fi f fi
k k d p p k e
! "
# $ = ! ! =
% &
. (4.3.12)
39


Figure 4.6 Contribution of diffusive flow to the apparent matrix permeability.

Figure 4.7 Error caused by neglecting diffusive flow in matrix for a 0.8 specific gravity gas.
40

In (4.3.12),
fi
k is the permeability at initial conditions and
f
d is a characteristic parameter
of the rock, which is determined experimentally (the range of
f
d is between
4
10
!
and
3
10
!
psi-1).
This definition corresponds to Type I rocks considered by Raghavan and Chin (2004). Other
relationships defining the stress-dependent permeability could also be used in the derivations
(Walsh, 1981, Gutierrez et al., 2000, Raghavan and Chin, 2004, Chipperfield et al., 2007).
Figure 4.8 shows the stress-dependent fracture permeability,
f
k , as a function of pressure
for
4
5 10
f
d
!
= " psi
-1
and 2000
fi
k = md at 5000
i
p = psi. It is shown in Fig. 4.8 that the first 1000
psi pressure drop causes 5% reduction in the initial permeability, and neglecting this reduction
causes 5.13% error in permeability at 4000 psi. In this example, a moderate value for
f
d is used,
which may be more appropriate for the reduction of permeability in rock matrix. For an
unpropped natural fracture in shale,
f
d should be much larger as indicated by Cho, 2011. Results
in Fig. 4.8 indicate that the reduction in natural fracture permeability due to stress dependency
may be significant and should be considered in modeling flow in fractured shale-gas reservoirs.

Figure 4.8 Stress-dependent fracture permeability and the error for neglecting the stress
dependency as a function of pressure.
41

4.3.3 Dual-Mechanism Dual-Porosity Reservoir Model and Transfer Function
As indicated by the results in Figs. 4.6 and 4.7, diffusive flow in low permeability shale
matrix may be significant and should be accounted for in modeling gas flow in source rocks. In
this section, diffusive flow in matrix is incorporated into a dual-porosity formulation for fractured
shale. Figure 4.9 shows the schematic of the dual-porosity system considered in the formulation.

Figure 4.9 Schematic of the system used to incorporate matrix diffusive flow into transient dual-
porosity model.
For convenience, in demonstrating the derivation of the dual-mechanism dual-porosity
reservoir model, radial flow in the reservoir (e.g., a fully penetrating vertical well) and spherical
geometry for the matrix blocks are considered. In the results section, however, the dual-
mechanism dual-porosity formulation developed here is applied to linear flow for a fractured
horizontal well in a source rock reservoir. Besides, the spherical matrix geometry chosen here is
not a requirement for our formulation; the formulation can be readily extended to other matrix
shapes. The details of the derivation of the dual-mechanism dual-porosity model incorporating
Darcy and diffusive matrix flows and stress-dependent natural fracture permeability are given in
Appendix C. Here, the special features and assumptions of the formulation are highlighted.
4.3.3.1 Assumptions
As is customary, the dual-mechanism dual-porosity model is derived in the Laplace
transform domain. Transient fluid flow is considered in the matrix system. To linearize gas-flow
equations, the pseudopressure approach is used. The pseudopressure is defined by
( )
0
2 ; or
p
g
p
m p dp m f
z
!
! ! !
" !

#
# = =
$
, (4.3.13)
where the subscripts m and f indicate matrix and fracture properties, respectively,
!
"
"
#
$%&!'(
)!%*&+!,
-,../0!,
"12
!1!
#
!
"
"
#
$%&!'(
)!%*&+!,
-,../0!,
"12
!1!
#
!
"
"
#
$%&!'(
)!%*&+!,
-,../0!,
"12
!1!
#
42

1
am
m
m
b
p
!
" #
$ %
= +
$ %
& '
, (4.3.14)
and
f f
d p
f
e !
" #
= . (4.3.15)
4.3.3.2 Matrix and Fracture Flow Equations
As shown in Appendix C.1, the matrix flow equation, incorporating Darcy and diffusive
flows, is given, in terms of pseudopressure, by
2
2
2 2
1 2 1
mD mD mD mD
D
D D D D D D mDi D
m m m m
r
r r r r r r t !
" #
$ $ $ $ $
% &
= + =
% &
$ $ $ $ $
' (
. (4.3.16)
As customary with the pseudopressure approach, the matrix flow equation is linearized by
assuming
( )
( )
f tf g ami
mi i
mD mDi
fi m tm g fi
i
c k
c k
!
"
" "
" !
# = = . (4.3.17)
The solution of the matrix flow equation is obtained in Appendix C.1 and given, in the Laplace
transform domain, by
( )
( )
( )
( ) ( )
sinh
, , 1 , ,
sinh
mD mDi D
mD D mD p mD fD mD
D mDi mD
r s r
m r R s f R s m R s
r s r
!
!
" # = $
% &
(4.3.18)
where
( )
( )
( )
,
,
,
bD mD
p mD
fD mD
m R s
f R s
m R s
= (4.3.19)
and
( ) ( ) ( ) , , , ,
bD mD fD mD mD D mD mD
m R s m R s m r r R s = ! = . (4.3.20)
In (4.3.20),
bD
m is a dimensionless pseudopressure defined by (4.1.2) with

43

( )
( )
( ) ( )
,
, 2 1 1
, , ,
i
f f
f m
p
d p d p
am
b m
p R t
g
f m m m m
b p
m R t e e dp
p z
m R t m r r R t

! ! " # "
$ % & '
!
( ) *+
! " = # +
( ) *+
!
, - . /
= " # " =
0
. (4.3.21)
The appearance of ( )
p
f s in the solution is due to the use of two different
pseudopressures for the matrix and fracture. The definition of pseudopressure includes the
pressure-dependent component of the permeability for each medium as shown in (4.3.14) through
(4.3.15). Appendix C.1 shows that this definition creates an inequality of pseudopressures at the
matrix-fracture interface even though the pressures are continuous at the interface. The difference
between the fracture and matrix pseudopressures is given by
b
m !

in (4.3.21). As shown in
Appendix C.2, the fracture flow equation, in terms of pseudopressure, is given by
( ) , ,
1 1
2
D mD mD D
fD fD
mD
D
D D D D fD D
r r R t
m m
m
R
R R R r t
!
"
=
# $ # $
% %
% %
& ' & '
( =
& ' & '
% % % %
) * ) *
!
, (4.3.22)
where
m
fi f
k L
k h
! =
!
. (4.3.23)
Assuming
1
fi
fD fDi
fi
!
! !
!
" = = , (4.3.24)
and in Laplace domain (4.3.22) yields the following equation
( ) , ,
1
2 0
D mD mD
fD
mD
D fD
D D D D
r r R s
m
m
R sm
R R R r
!
=
" # " #
$
$ $
% & % &
' ' =
% & % &
$ $ $
( ) ( )
!
. (4.3.25)
When the matrix solution (4.3.18) is used for the second term on the left hand side of (4.3.25),
then
( )
1
0
fD
D fD
D D D
m
R sf s m
R R R
! "
#
#
$ %
& =
$ %
# #
' (
. (4.3.26)
In (4.3.26),
( ) f s is the transfer function between the matrix and fractures given by
44

( ) ( )
15 15
, 1 1 1 coth 1 ,
5
ami
mD p mD
i
b s s
f R s f R s
s p
! " "
! !
# $
% & # $
' (
% & ) ' (* = + + + +
, -
' ( ' (
) *
. / , -
. /

(4.3.27)
where
( )
( ) ( )
( ) ( )
( ) ( )
( )
( )
3
2
4 3 2
2 4 2 3
t m t m t m
m m m
t f t m f t f
f f f
c V c r c r
c V c r h c h
! ! " !
#
! ! " !
= = =

(4.3.28)
and
( )
( )
( )
3
2 2 2
2
4 3 2
2 4 2 3
ami m ami m ami m
fi f fi m f fi f
k r k V k r
L L L
k V k r h k h
!
" # # #
!
$ % $ % & '
( ) ( ) * +
= = =
( ) ( ) * +
* +
( ) ( )
, - . / . /

(4.3.29)
The transfer function defined in (4.3.27), together with the definitions of ! and ! and the
definition of the matrix and fracture pseudopressures in (4.3.13) through (4.3.15) constitute the
main result of the dual-mechanism dual-porosity formulation for fractured source rock reservoirs.
Substituting these definitions for their counterparts in the conventional dual-porosity solutions
yields the solutions for dual-mechanism dual-porosity systems.
The numerical evaluation of the solutions with ( ) f s defined in (4.3.27) poses problems
because of the nonlinearity caused by the pressure dependence of ( )
p
f s defined in (4.3.19). As
noted above, ( )
p
f s is related to the dimensionless pseudopressure difference at the matrix-
fracture interface by the following relation:
( ) ( ) ( )
( ) ( )
, , 1 , , , ,
, , , ,
mD mD mD p mD mD fD mD mD
fD mD mD bD mD mD
m r R s f r R s m r R s
m r R s m r R s
! " = #
$ %
= #
. (4.3.30)
In general, it is also possible to express the pseudopressure difference at the interface in terms of
the flux as follows (Raghavan and Ozkan, 2011):
( ) ( ) , , , ,
D mD
fD
fD mD mD mD mD mD
D
r r
m
m r R s m r R s
r
=
! "
#
$ %
& = '
$ %
#
( )

(4.3.31)
where ! denotes the contact resistance of the interface [e.g., 0 ! = for perfect contact and
(4.3.31) becomes equivalent to pseudopressure continuity at the interface]. From (4.3.19),
(4.3.30), and (4.3.31), one can write
45

( )
( )
( ) ( )
, ,
, ,
, , , ,
D mD
fD
D
r r
bD mD mD
p mD mD
fD mD mD fD mD mD
m
r
m r R s
f r R s
m r R s m r R s
=
! "
#
$ %
$ %
#
& '
= = ( . (4.3.32)
Considering the zero storativity of the interface, it is reasonable to assume that the flux to
pseudopressure ratio represented by the last term in (4.3.32) is at steady-state; that is, independent
of time (this is akin to the assumption of steady-state thin skin condition at the surface of a
wellbore). This discussion leads to the iterative approach used in this work to evaluate the
solutions including a pressure-dependent ( ) f s . In general, the iterative approach requires the
evaluation of the solution first with 1
p
f = . In the subsequent iterations, the computed pressures
are used to calculate a new
p
f and the procedure is repeated until convergence is achieved within
a set tolerance.


46

CHAPTER 5
DISCUSSION OF RESULTS
This chapter presents the results obtained with the models developed in the previous
chapters. First the results on the dual-porosity model with matrix, microfracture and
macrofracture media are presented. Then, some results are discussed for the dual-porosity model
with Knudsen diffusive (slip) flow and stress-dependent fracture permeability reduction in
addition to Darcy flow.
5.1 Dual-Porosity Model with Matrix, Microfracture and Macrofracture Media
In this section, first the microfracture model will be used with synthetic data and the
results will be discussed. Later, the data of two wells in Haynesville and Barnett shale-gas
reservoirs will be discussed.
For the results presented based on the synthetic data (Figs. 5.1 through 5.5), the example
set of data given in Table 5.1 is used. The data have been selected by considering the ranges of
the properties of the US source-rock gas fields. The properties of the macrofractures and
microfractures, however, are only assumptions as these intrinsic properties are not measured (also,
the effective or bulk properties inferred from the well performance data depend on the
interpretation model). A ratio of 100:1 has been used for the macrofracture to microfracture
permeability and thickness. Similarly, it is assumed that the intrinsic porosity of the
microfractures is about 55% of that of the macrofractures. Finally, to highlight the effect of the
matrix microfractures, the matrix permeability is selected in the lower range. This selection is
also necessary to comply with the assumption of dual-porosity behavior in the matrix surface-
layer (negligible flow capacity in the matrix compared to microfractures) used in the derivation of
the solution. Because for very tight matrix permeability, the outer reservoir beyond the stimulated
reservoir volume does not considerably contribute to well performance for practical ranges of
time, the drainage area of the well is limited to that of the stimulated reservoir volume.
Additionally, spherical matrix radius is between 3 and 6 ft, macrofracture thickness is in the range
of 5x10
-4
to 5x10
-3
ft, and the microfracture thickness is 5x10
-6
ft. Microfractured surface layer
thickness is 20% of the matrix-block radius in the first data set (Table 5.1) while it is 4% of the
matrix-block radius in the field cases that are considered (Tables 5.2 and 5.3). These are, however,
rather arbitrary choices as no measurement is currently available and will be used to qualitatively
discuss the effect of microfractures. In addition, same intrinsic compressibility is used for matrix
and macro and microfractures. Bulk compressibility values in return were significantly lower for
47

macro and micro fractures. Variable compressibility scenarios for matrix and macro and micro
fractures were not considered and could be the subject of further studies.
Fig. 5.1 shows a comparison of the performances of the fractured horizontal well in
homogenous and fractured reservoirs, whose properties are given in Table 5.1, for a constant
bottomhole pressure of 500 psi
wf
p = . Two cases of fractured reservoirs are considered in Fig. 5.1;
a fractured reservoir without microfractures and a fractured reservoir with matrix microfractures.
The comparison of the results in Fig. 5.1 for the homogenous reservoir case and the dual-porosity
system without microfractures displays significant contribution of macrofractures to the
production performance of the well. Also, examination of the results for the two dual-porosity
reservoir cases with and without microfractures reveals the contribution of matrix microfractures
to the production performance of the well. This contribution becomes more evident and helps
sustain higher production rates at intermediate times. It should be noted, however, that the
accelerated production at early- and intermediate-times ends up in a more rapid decline at later
times. Another observation is that the cumulative gas production for the homogenous case has not
plateaued after thousand days and the macro and microfractured cases have already produced
most of the available gas by this time. Because a closed drainage volume is considered in this
figure, at the limit of zero production rate, the cumulative production for all three cases
considered in the figure will be the same. Because the existence of fractures accelerates
production at earlier times, these cases experience a steeper decline at later times.
In Fig. 5.1, the production behavior of the horizontal well for different thicknesses of the
matrix surface layer is considered. As explained earlier, the thickness of the matrix surface layer,
ms
h , depends on the penetration of the surface-connected microfractures into the matrix. The case
for 0
ms
h = corresponds to a homogenous (unfractured) matrix. (The results for this case have
been computed by using the model developed in this research with
mc m
r r = and also by using the
conventional dual-porosity model without microfractures. The two sets of the results were
identical and provided a means of verification for the model developed in this work.) Figure 5.2
shows that deeper penetration of surface-connected microfractures into the matrix (that is,
increased thickness of the microfractured surface layer), increases the contribution of the matrix
to the production performance of the well at intermediate times.


48

Table 5.1 Synthetic Data Used for Results Presented in Figs. 5.1 through 5.5

Formation thickness, h, ft 250
Macrofracture permeability , k
f
,
md
2,000
Wellbore radius, r
w
, ft 0.25 Macrofracture porosity, !
f
0.45
Horizontal well length, L
h
, ft 5,000
Macrofracture total
compressibility, c
tf
, psi
-1

2.5E-04
Distance to boundary parallel to
well, x
e
,

ft
250 Macrofracture thickness, h
f
, ft 5.0E-04
Number of hydraulic fractures 10 Macrofracture spacing, ft 12.5
Distance between hydraulic
fractures, 2y
e
, ft
500 Hydraulic fracture porosity, !
F
0.38
Viscosity, , cp 0.0184
Hydraulic fracture permeability,
k
F
, md
100,000
Matrix permeability, k
m
, md 1.0E-06
Hydraulic fracture total
compress, c
tF
, psi
-1

2.5E-04
Matrix porosity, !
m
0.05
Hydraulic fracture half-length, x
F
,
ft
250
Matrix total compressibility, c
tm
,
psi
-1

2.5E-04 Hydraulic fracture width, w
F
, ft 0.01
Microfracture permeability, k
mf
,
md
20 Initial reservoir pressure, p
i
, psi 3,800
Microfracture porosity, !
mf
0.25
Constant surface flow rate, q
sc
,
Mscf/d
940
Microfracture total compressibility,
c
tmf
, psi
-1

2.5E-04
Constant flowing bottomhole
pressure, p
wf
, psi
500
Microfracture thickness, h
mf
, ft 5.0E-06 Reservoir temperature, T, F 200
Microfracture spacing, ft 0.5 Molecular weight, M
g
, lb
m
/lb
m
-mol 16
Surface-layer thickness (microfrac
length), h
ms
, ft
0.2r
m
Specific gravity, SG 0.6
Matrix block radius, r
m
, ft 6.25

49


A Flow Rate versus Time

B Cumulative Production versus Time
Figure 5.1 Comparison of production performances of a fractured horizontal well in
homogeneous reservoir and dual-porosity reservoirs with and without microfractures.
As discussed earlier, the effect of matrix microfractures is expected to be more important
for tighter matrix permeabilities. Fig. 5.3 highlights the effect of matrix permeability on the
relative significance of matrix microfractures. For the case considered in Fig. 5.3, the effect of
microfractures decreases considerably when the matrix permeability increases from
6
10 md
m
k
!
=
50

to
4
10 md
!
. The effect of microfractures rapidly diminishes with further increase in matrix
permeability.
It is interesting to examine the effect of microfracture permeability on the production
performance of the well. In Fig. 5.4, three different permeabilities of the microfractures were
considered. Interestingly, the effect of microfracture permeability is only evident at very early
times. This may be explained by the limited supply of gas from the matrix into the microfractures
due to very low flow capacity of the matrix; in other words, there is not enough gas to flow in
microfractures.
Finally, the effect of microfractures on pseudopressure responses of multiple fractured
horizontal wells is considered. For the case considered in Table 5.1, with a constant surface
production rate of 940 Mscf/D, Fig. 5.5 shows the pseudopressure and derivative responses. The
figure clearly indicates the effect of microfractures on pressure-transient characteristics of the
well. The most important result to note from Fig. 5.5 is that the conventional flow regimes
indicated by certain derivative (or straight line) characteristics may not be identified when
microfractures contribute to the flow performance of the well. Also, the pressure-transient
characteristics displayed in Fig. 5.5 are different from those for triple-porosity systems because
the microfractures do not form a connected network (a continuum) to constitute the third porosity
in the system.
Additionally, two shale-gas field examples are considered to further verify the use of
microfractures in the tri-linear flow model. These examples included variable rate and
corresponding pressure data over time. The following superposition equation is used to generate
variable-rate pressure responses by using constant rate solution.
1 0
1
1, 422
( ) ( , ), 0
n
f i i fD D D Di
i
fi ft
T
m q q m R t with q
k h
!
"
=
# = " " =
$
(5.1.1)
In (5.1.1),
fD
m is the constant-rate dimensionless pseudopressure response obtained from
the trilinear model with and without the microfractures. In both models the objective is not to
achieve the best history match but to demonstrate the impact of microfractures given some set of
reservoir properties that reasonably matched the production history. In both cases, the same set
of microfracture properties as in Table 5.1 are used. First the production history is matched with
the microfractures included in the model; then, the same solution is generated by taking the
microfractures out of the model to gauge the impact of them.
51


A Flow Rate versus Time

B Cumulative Production versus Time
Figure 5.2 Effect of matrix surface layer thickness (microfracture penetration into matrix)
on production performances of a fractured horizontal well in a dual-porosity reservoir with
microfractures.
52


A Flow Rate versus Time

B Cumulative Production versus Time
Figure 5.3 Effect of matrix permeability on the impact of microfractures on production
performances of a fractured horizontal well.

53



A Flow Rate versus Time

B Cumulative Production versus Time
Figure 5.4 Effect of microfracture permeability on production performance of a fractured
horizontal well.
54


Figure 5.5 Effect of microfractures on pseudopressure responses of a fractured horizontal well.
The first example is from a horizontal shale-gas well in the Haynesville Formation. The
well production data presented in Fig. 5.6 are taken from Wang and Liu (2011). This well has a
depth of 10,600 ft and its lateral length is 3,200 ft. It was completed with 10 stages and has a
production history of more than three months. Most of the effort was concentrated on history
matching the production after the initial shut-in period.
The data presented in Table 5.2 are used to match the production performance
(microfracture properties were kept the same as those in Table 5.1). General reservoir properties,
including formation thickness, temperature, and matrix porosity, are in the ranges that are
reported by Wang and Liu (2011). Matrix permeability and macrofracture properties were varied
to match the well performance after the initial shut-in period. Because the focus was to match the
later portion of the production history, the model was initialized at a much lower initial reservoir
pressure (3,800 psia) rather than using the reported 9,000 psia pressure range (up to 0.9 psi/ft
pressure gradient).
In Fig. 5.7, the bottom hole pressure history match with and without the microfractures is
presented. As explained in the discussion of the data, the focus of the history match was the
pressure data after the initial shut-in period. First, the pressure history was matched with matrix
microfractures in the trilinear model; then keeping the other match parameters fixed,
55

microfractures were removed from the model and another set of pressure responses were
generated. The comparison of the two sets of the generated pressure responses in Fig. 5.7
highlights the impact of microfractures on matching the well performance. Having microfractures
in the model helps to match the pressure responses even in the first 100 days, which indicates
earlier matrix support due to matrix-surface stimulation caused by microfractures. It should be
noted that a match was achieved with relatively low matrix permeability (
6
3.2 10 md
m
k x
!
= ).
Other attempts (not shown here) to match the production performance without matrix
microfractures yielded much higher matrix permeability (and the match was not as good). This
result is relevant to the discussion of the discrepancies between the estimates of shale-matrix
permeability from core measurements and actual production performance (Cluff et al., 2007).


Figure 5.6 Production history of a fractured, horizontal gas well in the Haynesville Shale (Wang
and Liu, 2011).

The second example considered is from a horizontal gas well in the Barnett Shale. The
well production data presented in Fig. 5.8 is reproduced from Anderson at al. (2010). They did
not report reservoir properties and well specific data so the data given in Table 3 consist of
assumed values. A well that had 3600 ft lateral length which was completed in 12 stages was
considered. The well has a production history of more than three years and it was attempted to
history match the initial 600 days.
56

The data presented in Table 5.3 was used to match the production performance. As in the
Haynesville well considered above, the microfracture properties were kept fixed (the same as
those used in the Table 5.1) and matrix permeability and macrofracture properties were varied to
match the well performance.
Table 5.2 Data used for Matching the Haynesville Shale-Gas Well Performance

Formation thickness, h, ft 270
Macrofracture permeability , k
f
,
md
300
Wellbore radius, r
w
, ft 0.25 Macrofracture porosity, !
f
0.45
Horizontal well length, L
h
, ft 3200
Macrofracture total
compressibility, c
tf
, psi
-1

2.5E-04
Distance to boundary parallel to
well, x
e
, ft
500 Macrofracture thickness, h
f
, ft 1.0E-03
Number of hydraulic fractures 10 Macrofracture spacing, ft 10
Distance between hydraulic
fractures, 2y
e
, ft
320 Hydraulic fracture porosity, !
F
0.389
Viscosity, , cp 0.0184
Hydraulic fracture permeability,
k
F
, md
100,000
Matrix permeability, k
m
, md 3.2E-06
Hydraulic fracture total
compressibility, c
tF
, psi
-1

2.5E-04
Matrix porosity, !
m
0.07 Hydraulic fracture half-length, x
F
, ft 200
Matrix total compressibility, c
tm
, psi-
1
2.5E-04 Hydraulic fracture width, w
F
, ft 0.01
Microfracture permeability, k
mf
, md 20 Initial reservoir pressure, p
i
, psi 3,800
Microfracture porosity, !
mf
0.25 Reservoir temperature, T, F 290
Microfracture total compressibility,
c
tmf
, psi-1
2.5E-04 Molecular weight, Mg, lb
m
/lb
m
-mol 16
Microfracture thickness, h
mf
, ft 5.0E-06 Specific gravity, SG 0.7
Microfracture spacing, ft 0.5 Matrix block radius, r
m
, ft 5.0
Surface-layer thickness
(microfracture length), h
ms
, ft
0.04r
m

The pressure match of the Barnett field example is given in Fig. 5.9. As indicated by the
flow rate data in Fig. 5.8, between 58 and 71 days, the well was shut-in but the corresponding
pressure data did not display pressure build-up characteristics. Because of this inconsistency, the
portion of the data between 58 and 71 days was ignored in the pressure match shown in Fig. 5.9.
It can be seen that even after 600 days of production, there is no clear separation between the
pressures predicted by the models with and without microfractures. Although this result may be
attributed to the lack of matrix microfractures in this case, an alternate explanation is provided
below.

57


Figure 5.7 Pressure match of the Haynesville shale gas reservoir well with/without microfracs
(Matrix Permeability; 3.2E-06 md).



Figure 5.8 Production history of a horizontal gas well in the Barnett Shale (Anderson, 2010).



58

Table 5.3 Data used for Matching the Barnett Shale-Gas Well Performance

Formation thickness, h, ft 120 Macrofracture permeability , k
f
, md 300
Wellbore radius, r
w
, ft 0.23 Macrofracture porosity, !
f
0.45
Horizontal well length, L
h
, ft 3600
Macrofracture total
compressibility, c
tf
, psi-1
2.5E-04
Distance to boundary parallel to
well, x
e
, ft
105 Macrofracture thickness, h
f
, ft 5.0E-03
Number of hydraulic fractures 12 Macrofracture spacing, ft 6
Distance between hydraulic
fractures, 2y
e
, ft
300 Hydraulic fracture porosity, !
F
0.38
Viscosity, , cp 0.0184
Hydraulic fracture permeability, k
F
,
md
100,000
Matrix permeability, k
m
, md 2.9E-04
Hydraulic fracture total
compressibility, c
tF
, psi-1
3.17E-04
Matrix porosity, !
m
0.036 Hydraulic fracture half-length, x
F
, ft 100
Matrix total compressibility, c
tm
,
psi
-1

2.5E-04 Hydraulic fracture width, w
F
, ft 0.001
Microfracture permeability, k
mf
,
md
20 Initial reservoir pressure, p
i
, psi 2,400
Microfracture porosity, !
mf
0.25 Reservoir temperature, T, F 106
Microfracture total
compressibility, c
tmf
, psi
-1

2.5E-04 Molecular weight, Mg, lb
m
/lb
m
-mol 16
Microfracture thickness, h
mf
, ft 5.0E-06 Specific gravity, SG 0.8
Microfracture spacing, ft 0.5 Matrix block radius, r
m
, ft 3.0
Surface-layer thickness
(microfracture length), h
ms
, ft
0.04r
m


As noted earlier, the same microfracture properties were used in both field examples
(Tables 5.2 and 5.3). Macrofracture properties are all the same also except for the macrofracture
spacing (10 ft for Haynesville vs. 6 ft for Barnett) and macrofracture thicknesses (1x10
-3
ft for
Haynesville vs. 5x10
-3
ft for Barnett). These differences are not significant enough to explain why
the Haynesville example exhibits considerable microfracture contribution while the Barnett
example does not. The major difference between the two examples is the matrix permeability;
there is two orders of magnitude difference between the matrix permeabilities of the two cases
and the lower matrix permeability case (Haynesville example with a matrix permeability of
3.2x10
-6
md) shows a more convincing microfracture impact. This is also consistent with the
earlier results that were discussed in Fig. 5.3 where the microfracture impact is shown to diminish
quickly when the matrix permeability reaches about 10-4 md.
59


Figure 5.9 Pressure match of the Barnett shale gas reservoir well with/without microfracs
(Matrix Permeability; 2.9E-04 md).

5.2 Dual-Porosity Model with Knudsen Diffusion (Slip Flow) in Matrix and Stress-
Dependent Fracture Permeability
In this section, a multiply fractured horizontal well in a source-rock gas reservoir is
considered and the dual-mechanism dual-porosity approach introduced in Section 4.3 is applied.
The trilinear flow model of Ozkan et al. (2009) and Brown et al. (2009) is used to represent flow
in the reservoir. The trilinear model considers three mutually orthogonal flow regions; one in
finite-conductivity hydraulic fractures, one in the naturally fractured zone between hydraulic
fractures (the stimulated reservoir volume, SRV, around the well and the hydraulic fractures), and
one in the outer reservoir beyond the tips of the hydraulic fractures (made of unfractured source
rock). Hydraulic fractures are assumed to be identical and uniformly distributed along the length
of the horizontal well. Therefore, each hydraulic fracture produces 1
F
n of the total flow rate,
where
F
n is the number of hydraulic fractures. In the examples below, the data given in Table 5.4
is used, unless otherwise stated.
5.2.1 Comparison of the Dual-Porosity Formulations with Spherical Matrix Blocks and Slabs
A multiply fractured horizontal well in a source-rock gas reservoir is considered with the
properties given in Table 5.4. To verify the dual-porosity formulation of these derivations with
spherical matrix blocks, the pseudopressure and derivative responses for constant matrix and
fracture permeabilities are computed and the results are compared to those of a conventional
60

transient dual-porosity model with slab matrices (de Swaan O, 1976, Kazemi, 1969, Serra et al.,
1983). The trilinear model for the multiply fractured horizontal well considered here is given by
Brown et al. (2009) for slab matrices. To obtain the solution for spherical matrix blocks, the
substitution of the dual-porosity parameters and the transfer function defined in (4.3.27), (4.3.28),
and (4.3.29) in the solution given by Brown et al. (2009) is made. Also, for a fair comparison,
adjustments are made to equate the matrix to fracture volume ratio in a representative volume
element in both models. Fig. 5.10 shows that the agreement between the two sets of the results is
excellent. This verifies the dual-porosity formulation with spherical matrix blocks used in this
research.
Table 5.4 Data Used for Results

Specific gravity of gas, SG 0.55 Initial fracture permeability, k
fi
, md 2000
Molecular weight of gas, M, lb
m
/lb
m
-mol 16 Initial fracture porosity, !!" 0.45
Initial reservoir pressure, p
i
, psi 2300
Initial fracture compressibility, c
tfi
,
psi
-1

9.E-04
Reservoir temperature, T,
o
F 109 Fracture thickness, h
f
, ft 0.001
Formation thickness, h, ft 250 Number of fractures in pay thickness 20
Wellbore radius, r
w
, ft 0.25 Hydraulic fracture porosity, !# 0.38
Reservoir size perpendicular to well
axis, x
e
, ft
500
Hydraulic fracture permeability, k
F
,
md
1.E+05
Half-distance between hydraulic
fractures, y
e
, ft
250
Hydraulic fracture compressibility,
c
tF
, psi
-1

9.E-04
Initial viscosity,
"
, cp 0.0184 Hydraulic fracture half-length, x
F
, ft 250
Constant matrix permeability, k
m
, md 1.E-08 Hydraulic fracture width, w
F
, ft 0.01
Initial matrix porosity, !$" 0.05
Hydraulic fracture flow rate, q
F
,
Mscf/D
94
Initial matrix compressibility, c
tmi
, psi
-1
9.E-04

5.2.2 Production from Source Matrix
The contribution of matrix flow is essential for the sustained productivity of source rock
gas wells. Here, flow in a spherical matrix block adjacent to a hydraulic fracture of a multiply
fractured horizontal well is considered. The schematic of the system considered in this example is
shown in Fig. 5.11 and the other relevant data are given in Table 5.4.

61



Figure 5.10 Comparison of pseudopressure and derivative responses for dual-porosity models
with slab and spherical matrix blocks.

Figure 5.11 Schematic of a spherical shale-matrix block adjacent to a horizontal well fracture.

1.E+04
1.E+05
1.E+06
1.E+07
1.E+08
1.E+09
1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04
TIME, Days
P
S
E
U
D
O
P
R
E
S
S
U
R
E

D
R
O
P
,

!
m
w
f
,

p
s
i
2
/
c
p
L
O
G
A
R
I
T
H
M
I
C

D
E
R
I
V
A
T
I
V
E
,

p
s
i
2
/
c
p
SLABS PSEUDOPRESSURE
SLABS DERIVATIVE
SPHERES PSEUDOPRESSURE
SPHERES DERIVATIVE
!
"
"
#
$%&'()
*'%+&,'-
.-//01'-
"23
424
#
!2!
#
567'%,/(+
*'%+&,'-
4
!
"
"
#
$%&'()
*'%+&,'-
.-//01'-
"23
424
#
!2!
#
567'%,/(+
*'%+&,'-
4
62

Fig. 5.12 shows the dimensionless matrix flow rate profile given by (5.2.1) crossing the
dimensionless radius
D
r in the spherical matrix block after 27.3 years of production.
( ) ( )
( ) , ,
, ,
mD D D D f msc sc
r R t
q r R t n q q = , (5.2.1)
where
msc
q is the volumetric flow rate, at standard conditions, crossing a radius r in the matrix
block and
f
n is the number of natural fractures in the pay thickness,
In Figure 5.12, only Darcy flow is considered in the matrix. The flow rate profiles in
Figure 5.12 indicates that, for permeabilities close to nanodarcy range, most of the matrix
production comes from the volume immediately below the surface of the matrix block even after
27.3 years of production. In other words, the core of the matrix block is not drained. Although not
shown here, for matrix blocks farther away from the hydraulic fracture, production is limited to a
thin layer on the surface of the matrix only. The results shown in Figure 5.12 are one of the
motivations of this study to incorporate additional mechanisms of flow in matrix.

Figure 5.12 Matrix flow rate profiles after 27.3 years of production.

5.2.3 Effect of Slip Flow in Matrix and Stress-Dependent Permeability on the Performances of
Fractured Horizontal Wells
Fig. 5.13 shows the combined effect of slip flow in matrix and stress-dependent
permeability in fractures on pseudopressure and derivative responses of a multiply fractured
horizontal well in a shale-gas reservoir. The data given in Table 5.4 have been used for this figure.
0.0000
0.0002
0.0004
0.0006
0.0008
0.0010
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
DIMENSIONLESS DISTANCE FROM MATRIX CENTER, r
D
= r/r
m
D
I
M
E
N
S
I
O
N
L
E
S
S

F
L
O
W

R
A
T
E

I
N

M
A
T
R
I
X
q
m
D
=

n
f
x

(
q
m
s
c
/
q
s
c
)

k
m
= 1E-8 md
k
m
= 1E-7 md
k
m
= 1E-6 md
k
m
= 3E-6 md
k
m
>= 5E-6 md
Matrix flow rate profile after 27.3 years
63

Significant improvement in well performance (smaller bottomhole pseudopressure drop) is
observed as time progresses in Fig. 5.13 when the effect of slip flow and stress-dependent
fracture permeability is taken into consideration. (Note that, for the case without slip flow and
stress-dependent permeability, the well ceases production around 1500 hr because the bottomhole
pressure reaches the absolute minimum.)
Intuitively, the contribution of slip flow in matrix is expected to improve the performance
of the well and stress-dependent fracture permeability is thought to reduce well performance
when pressure drops in the system as production continues. With this interpretation, the results in
Fig. 5.13 also indicate that the contribution of slip flow surmounts the performance reduction
caused by the reduction in natural-fracture permeability. This conclusion, however, requires
scrutiny.

Figure 5.13 Combined effect of slip flow and stress-dependent natural fracture permeability on
pseudopressure and derivative responses.
It is important to note that the comparison in Fig. 5.13 may not represent the true
magnitude of the contributions of slip flow and stress-dependent permeability. Because of the
dual-porosity assumption, wellbore pseudopressures shown in Fig. 5.13 are defined based on
64

natural-fracture properties. Thus, the definitions of pseudopressures with and without the effect of
stress-dependent permeability are different. To alleviate this issue, Fig. 5.14 presents the
comparison of the results in terms of the wellbore pressure drop. Again, significant improvement
in well performance (smaller bottomhole pressure drop) is observed as time progresses. In fact, as
will be discussed below in Figs. 5.15 and 5.16, for the case considered in Figs. 5.13 and 5.14, the
effect of stress on natural-fracture permeability is negligible. Therefore, the definition of the
pseudopressure is not significantly influenced by the stress-dependency of natural fractures and
the comparison in terms of pseudopressures (Fig. 5.13) is meaningful for the most part.

Figure 5.14 Combined effect of slip flow and stress-dependent natural fracture permeability on
pressure responses.
In Fig. 5.15, the effect of stress-dependent fracture permeability is examined without slip
flow in matrix (that is, considering only Darcy flow in matrix). The figure indicates a deviation of
the results for the stress-dependent natural-fracture permeability starting around 50 hr and the
system ceases production around 1000 hr. However, when we examine the same results in terms
of wellbore pressure drop in Fig. 5.16, we can see that the effect of natural-fracture permeability
reduction is negligible for all cases with or without slip flow in matrix. This result can be
explained by noting that the first upward bend of the pressure responses without the effect of slip
flow and fracture stress dependency at around 10
-2
hr in Fig. 5.16 indicates the start of the
depletion of the fracture network. The pressure drop in the fractures at this time is around 2.2 psi
65

which does not cause noticeable change in fracture permeability. By 0.5 hr, the fluid stored in the
fracture network is mostly depleted and if the support of the matrix is slow (Darcy flow in the
matrix only), there is not much volume of fluid to be transmitted by the fractures. As a result, the
reduced permeability is still sufficient to maintain the flow of fluids to the wellbore. If the matrix
support is stronger due to slip flow contribution, then, as shown in Fig. 5.16, pressure drop in the
fracture network is much smaller, which, in turn, yields insignificant reduction in fracture
permeability. Again, at late times (after approximately 15 hr), the pressure drop becomes large
enough to considerably reduce the permeability, but then there is not much fluid to flow in the
fractures (due to the decline in the support of fractures).

Figure 5.15 Effect of stress-dependent fracture permeability on pseudopressure and derivative
responses.
66


Figure 5.16 Effect of stress-dependent fracture permeability on pressure responses.
67

CHAPTER 6
CONCLUSIONS AND RECOMMENDATIONS
In this research a more detailed description of flow in shale matrix has been established
and incorporated into modeling of flow in fractured shale reservoirs. In addition, fluid transfer
from shale matrix to a fracture network has been scrutinized and the coupling of interface
conditions have been improved. The research has focused on two important problems:
i. the effect of microfractures on fluid transfer from the shale matrix to the fracture network
and
ii. incorporating slip-flow in matrix nanopores and stress-dependent fracture permeability
into dual-porosity modeling of shale reservoirs.
The current findings and accomplishments of the research in these two areas are
summarized below. Additionally these two models are integrated in the recommendations section,
and solutions for a model with microfractures and slip-flow in matrix and stress-dependent
fracture permeability are presented for consideration in future studies.
6.1 Conclusions
6.1.1 Effect of (Matrix-Surface) Microfractures
An analytical model has been used with composite matrix blocks consisting of a
homogenous core, where unconnected microfractures do not considerably contribute to flow
capacity, and a matrix surface layer, where the microfractures connected to the matrix surface
(resembling wormholes) cause a stimulation effect. The composite matrix flow was then coupled
with the flow in a network of macrofractures as in the conventional dual-porosity idealizations of
fractured media. This model was used to investigate the effect of matrix-surface stimulation and
its contribution to improved fluid transfer from the matrix medium to the fracture network. This
work has demonstrated the importance of microfractures in shale, which would be usually
ignored in more conventional formations. The following remarks and conclusions are warranted
based on the research presented in this dissertation:
1) Microfractures in matrix may contribute up to 10 percent to the production performance
of multiply fractured horizontal wells in ultratight formations considered in this
dissertation. Therefore, the characterization of microfractures should be taken as an
important issue.
68

2) The significance of microfractures becomes less important when the order of magnitude
contrast between the microfracture and matrix permeabilities decreases (this research
indicates that a minimum of five orders of magnitude permeability contrast is required to
have a sizeable contribution of the microfractures).
3) Matrix microfractures improve fluid transfer from the matrix medium to the fracture
network and accelerate production by providing earlier and more effective contribution
of matrix.
4) The contribution of the matrix due to microfractures should not be incorporated into
conventional dual-porosity models in the form of enhanced matrix permeability as the
microfractured surface layer of the matrix causes different flow characteristics than a
homogeneous (unfractured) matrix.
5) The effect of the microfractured surface layer of the matrix should not be taken into
account by a triple-porosity model either. Triple-porosity modeling is appropriate to
incorporate two sets of connected natural fractures (both forming a continuum for
modeling purposes) or connected fractures and vugs. The microfractures observed in
shale samples are short, discontinuous, and not in a continuum. Except for the
microfractures located closer to the interface of the matrix and the macrofractures, they
do not provide a continuous flow medium.
6) The considerable effect of a microfractured matrix surface layer demonstrated in this
work indicates that matrix-surface stimulation may be a viable option for enhanced
hydrocarbon recovery from ultratight unconventional reservoirs.
6.1.2 Combined Effect of Slip-Flow and Stress-Dependent Permeability
The description of flow in a tight matrix has been improved in this research by
considering the diffusive (Knudsen) flow in nanopores. Furthermore, stress-dependent
permeability in the fracture network has been considered. The following is a short-list of the
accomplishments as a result of these efforts:
1) A dual-mechanism approach has been proposed. In dual-mechanism flow, when Darcy
flow becomes insignificant due to nanodarcy matrix permeability, Knudsen diffusion
(slip flow) takes over and contributes substantially to the transfer of fluids from matrix
to a fracture network.
2) By incorporating Darcy and diffusive flows in the matrix and stress-dependent
permeability in the fractures, a dual-mechanism dual-porosity formulation of flow in
fractured shale reservoirs has been developed.
69

3) In addition, a new dual-porosity transfer function for fractured shale-gas reservoirs has
been derived.
6.2 Recommendations for Future Work
During the last phase of this research, the two models incorporating the effect of matrix
microfractures and the effects of slip flow and stress-dependent natural-fracture permeability
developed in this research have been merged and the solution for an integrated model has been
derived. The objective of this effort was to demonstrate the potential of extensions of the
analytical approach presented in this dissertation. Specifically, the complications which could be
caused by the multi-level nonlinearity of the dual-porosity transfer function were of concern. This
effort was limited to the derivation of the integrated solution and the assessment of the
computational complications. Below, a summary of the integrated solution is presented with the
full details of the derivation in Appendix D.
For a cylindrical, dual-porosity reservoir with a fully penetrating vertical well, the
macrofracture flow equation in Laplace-transform domain for the integrated system is given by
1
0
fD
D fD
D D D
m
R um
R R R
! "
#
#
$ %
& =
$ %
# #
' (
, (6.2.1)
where
( ) u sf s =
, (6.2.2)
and
( ) 1
f m
f s f f ! = " . (6.2.3)
In (6.2.3),
( )
( )
, 3 1
1 tanh 1
3 ,
m m mfDi bsD D m
m
mfDi m mfDi mfD D
m r s
f s
s m r s
! " #
"
# ! #
$ % & ' ( )
* *
+ ,
- .
= / /
0 1
+ ,
- .
+ , * *
- .
2 3 4 5 6 7
, (6.2.4)
( ) ( )
( )
( ) ( )
( )
( )
( )
exp 2
,
1
5 exp 2 ,
m mf m mf m mD mcD
mf mD bD mD
f
m mm m mf m mf m mD mcD fD mD
u f u f u r r
h r m R s
f
u h u f u f u r r m R s
! "
# $ % &
# $
' ' + '
( ( ) *
+ , - .
= '
/ 0
+ , - .
( ( # $
' + + '
+ , 1 2 ) * ) * 3 4

(6.2.5)
and
70

( )
2
2
coth 1
12
mm m
mf mDi mcD mDi mcD
mcD
h
f s r s r
r L
!
" "
# $
= %
& '
. (6.2.6)
Other dimensionless variables used in (6.2.1) through (6.2.5) have been provided in the text and
in Appendix D. Most important for our discussion here are the terms denoted by
bsD
m and
bD
m ,
which correspond to a pseudopressure mismatch at matrix core and microfracture and surface-
layer and macrofracture interfaces.
As discussed earlier in this dissertation, the use of the transfer function given by (6.2.3)
incorporates the combined effects of slip flow, microfractures, and stress-dependent fracture
permeability. In Section 4.3.3.2, it was shown that the mismatch of the pseudopressure definitions
at the matrix-fracture interfaces requires an iterative solution procedure and poses some
computational problems. In this dissertation (Section 4.3.3.2), only one such condition was
considered between the shale matrix and the macrofractures. When matrix surface-layer
microfractures are considered with the assumption of stress-dependent fracture permeability, as in
the integrated solution discussed above, two coupling conditions between the matrix and
microfractures and microfractures and macrofractures give rise to interface conditions with
surface resistance. As a result, two boundary terms,
bsD
m and
bD
m , appear in the solution and
require an iterative approach for numerical evaluations. The nested structure of the functions
including the boundary terms is a warning sign for development of a stable iterative procedure.
Based on these findings, it is suggested that future studies to use analytical solutions integrating
multiple pressure-dependent permeability terms first consider the computational viability in light
of the discussion provide above.
71

NOMENCLATURE
Symbol Definition Unit
b Gas Slippage Factor psi
c Compressibility psi
-1
C Concentration lbm-mol/ ft
3

C
FD
Conductivity -
C
RD
Conductivity -
d Characteristic Parameter of Rock psi
-1
D Diffusitivity Coefficient ft
2
/D
f Transfer Function -
F Mass Influx lbm/ft
3
/D
h Thickness ft
k Permeability md
L Characteristic Length ft
m Pseudopressure psi
2
/cp
M Point in Reservoir -
M
g
Molecular Weight of Gas lbm/lbm-mol
n Number of Specific Features -
p Pressure psi
q Production Rate Mscf/D
Q Cumulative Production Mscf
r Radial Distance in Spherical Coordinates ft
R Radial Distance in Radial Coordinates ft
72

R
g
Gas Constant, 10.73 psi-ft
3
/lb
m
mol-
o
R
S Saturation -
s Laplace Transform Parameter -
s
c
Skin factor due to flow choking -
SG Specific Gravity -
T Reservoir Temperature
o
R
t Time hrs, D
v Volumetric velocity ft/D
v
rm
Radial matrix volumetric flow velocity ft/D
v
prm
Radial matrix Darcy flow velocity ft/D
v
srm
Radial matrix slip flow velocity ft/D
V Volume ft
3

w Width ft
x Length in direction perpendicular to hor well axis, ft
y Length in direction parallel to horizontal well axis ft
z Gas Compressibility Factor fraction

01$$2
!
!
Conversion constant = 2.637x10
-4
for t in hr and
6.328x10
-3
for t in days
! Partial derivative -
! Difference operator -
! Density lbm/ft
3

! Shape Factor ft
2

! Transmissivity -
73

! Porosity fraction
! Linear distance ft
! Diffusivity ft
2
/hr, ft
2
/D
! Viscosity cp
! Dummy variable -
! %3*).3'4'35 6

%789()':39
. ;::.)</3 6
= >.)(5 6
> >'?</9'*/@<99 6
& $A3<)/.@ '
! -.37).@ #).(37)<B&.()*C).(37)< 6
# D5=).7@'( #).(37)< 6
E 0.9 6
' F/'3'.@ 6
G D*)'H*/3.@ 6
? &.3)'A 6
?( &.3)'A (*)< 6
?C &'()*C).(37)< 6
?? %7)C.(< @.5<) ?.3)'A 6
?9 &.3)'A 97)C.(< @.5<) 6
I I73<) 6
) 1.='.@ 6
1 1<9<)4*') 6
() %3./=.)= +*/='3'*/ 6
* ,*3.@ 6
J K<@@8*)< 6
JC #@*J'/E K<@@8*)< 6


74

REFERENCES CITED
Abdassah, D. and Ershaghi, I. 1986. Triple-Porosity Systems for Representing Naturally
Fractured Reservoirs, SPE-13409-PA, SPE Form Eval pp. 113-127, April 1986.
Agarwal, R.G. 1979. Real Gas Pseudo-Time - A New Function for Pressure Buildup Analysis of
MHF Gas Wells, SPE 8279, SPE Annual Technical Conference and Exhibition, Sept. 23-26,
1979, Las Vegas, NV.
Al-Ahmadi A. H., Wattenbarger, R. A. 2011. Triple-Porosity Models: One Further Step Towards
Capturing Fractured Reservoirs Heterogeneity, Paper SPE 149054, presented at the SPE/DGS
Saudi Arabia Section Technical Symposium and Exhibition, 15-18 May 2011, Al-Khobar, Saudi
Arabia.
Al-Ghamdi, A., Ershaghi, I. 1996. Pressure Transient Analysis of Dually Fractured Reservoirs,
SPE 26959-PA, SPE J. March pp. 93-100, 1996.
Al-Hussainy, R., Ramey, H. J., and Crawford, P. B. 1966 The Flow of Real Gases through Porous
Media, JPT 18 (5): 624-636. SPE 1243-A.
Al-Hussainy, R., and Ramey, H. J. 1966. Application of Real Gas Theory to Well Testing and
Deliverability Forecasting, JPT 18 (5): 637-642. SPE 1243-PA.
Anderson, D. M., Nobakht M., Moghadam S., and Mattar L. 2010. Analysis of Production Data
from Fractured Shale Gas Wells, SPE 131787, SPE Unconventional Gas Conference, February
23-25, Pittsburgh, Pennsylvania.
Bello, R. O., 2009. Rate Transient Analysis in Shale Gas Reservoirs with Transient Linear
Behavior. Ph.D. Dissertation, Texas A & M University, College Station, Texas.
Bello, R. O., and Wattenbarger, R. A. 2008. Rate Transient Analysis in Shale Gas Reservoirs,
Paper SPE 114591, presented at the CIPC/SPE Gas Technology Symposium 2008 Joint
Conference, 16-19 June 2008, Calgary, Alberta, Canada.
Bello, R. O., and Wattenbarger, R. A. 2009. Modeling and Analysis of Shale Gas Production with
a Skin Effect, Paper CIPC 2009-082, presented at the Canadian International Petroleum
Conference, 16-18 June 2009, Calgary, Alberta, Canada.
Bello, R. O., and Wattenbarger, R. A. 2009. Multistage Hydraulically Fractured Shale Gas Rate
Transient Analysis, paper SPE 126754, presented at the SPE North Africa Technical Conference
and Exhibition, 14-17 February 2009. Cairo, Egypt,.
Barenblatt, G. E., Zheltov, I. P., and Kochina, I. N. 1960. Basic Concepts in the Theory of
Homogeneous Liquids in Fissured Rocks, Jour. Appl. Math. Mech., 24, No. 5, 1286--1303.
Beskok, A., Micro Flow. http://www.cfm.brown.edu/people/beskok/home.html
Best, M.E., and Katsube, T.J. 1995. Shale Permeability and its Significance in Hydrocarbon
Exploration, Geological Survey of Canada 14(3): 165-170
Bowker, K. A., 2007b. Development of the Barnett Shale Play, Fort Worth Basin: Search and
Discovery Article #10126:
http://searchanddiscovery.net/documents/2007/07023bowker/index.htm, accessed June 2009.
75

Brown, M., Ozkan, E., Raghavan, R., and Kazemi, H. 2009. Practical Solutions for Pressure
Transient Responses of Fractured Horizontal Wells in Unconventional Reservoirs, Paper SPE
125043 presented at the SPE Annual Technical Conference and Exhibition, 4-7 October, New
Orleans, Louisiana.
Bruner, K. R., Smosna R. 2011. A Comparative Study of the Mississippian Barnett Shale, Fort
Worth Basin, and Devonian Marcellus Shale, Appalachian Basin. DOE/NETL-2011/1478, April
2011.
Carslaw, H. S., and Jaeger, J. C. 1959. Conduction of Heat in Solids, Second ed. Oxford
University Press, Oxford.
Celis, V., Guerra, J., and Prat, G. Da, 1994, A New Model for Pressure Transient Analysis in
Stress Sensitive Naturally Fractured Reservoirs, SPE Advanced Technology Series 2 (1): 126-135.
Chen, L. Y., Raghavan, R., and Thomas, L. K. 2000. Fully Coupled Analysis of Well Responses
in Stress-Sensitive Reservoirs, Paper SPE 66222 presented at SPE Annual Technical Conference
and Exhibition, 27-30 September, New Orleans, Louisiana.
Chen, L. Y., Raghavan, R., and Thomas, L. K. 2000. Fully Coupled Geomechanics and Fluid-
Flow Analysis of Wells With Stress-Dependent Permeability, Paper SPE 58968 presented SPE
International Conference and Exhibition, 2-6 November, Beijing, China.
Chipperfield, S. T., Wong, J. R., Warner, D. S., Cipolla, C. L., Mayerhofer, M. J., Lolon, E. P.,
and Warpinski, N. R. 2007. Shear Dilation Diagnostics: A New Approach for Evaluating Tight
Gas Stimulation Treatments, Paper SPE 106289 presented SPE Hydraulic Fracturing Technology
Conference, 29-31 January, College Station, Texas.
Cho, Y. 2011. Effects of Pressure-Dependent Natural-Fracture Permeability on Shale-Gas Well
Production, M. Sc. Thesis, Colorado School of Mines, Golden, Colorado
Cluff, R. M., Shanley, K. W., and Miller, M. A. 2007. Three things we thought we understood
about shale gas, but were afraid to ask..., poster presentation, AAPG 2007 Annual Conference,
April 1-4, Long Beach, CA.
Dreier, J. 2004. Pressure-Transient Analysis of Wells in Reservoirs with a Multiple Fracture
Network, M.Sc. Thesis, Colorado School of Mines, Golden, Colorado.
de Swaan-O, A. 1976. Analytic Solutions for Determining Naturally Fractured Reservoir
Properties by Well Testing, SPEJ 16 (3): 117-122. SPE 5346-PA
El-Banbi, A. H. 1998. Analysis of Tight Gas Wells, Ph.D Dissertation, Texas A & M University,
College Station, Texas.
Engelder, T., Lash, G. G. 2008. Marcellus Shale Plays Vast Resource Potential Creating Stir in
Appalachia, The American Oil and Gas Reporter, May 2008, pp. 7
Ertekin, T., King, G. R., and Schwerer, F. C. (1986). Dynamic Gas Slippage: A Unique Dual-
Mechanism Approach to the Flow of Gas in Tight Formations, SPE Formation Evaluation (Feb.)
pp. 43-52.
Fred F. Meissner, Basin Center Hydrocarbon Accumulations, Examples from Rocky Mountain
Basins,Analogs for the World, RMAG Basin Center Gas Symposium, 2007,
76

Gringarten, A., and Ramey, H. 1973. The Use of Source and Greens Functions in Solving
Unsteady-Flow Problems in Reservoirs, SPE 3818-PA, SPEJ OCT 1973.
Guegen, Y., Sarout, J., Fortin, J. and Schubnel A., Cracks in Porous Rocks: Tiny Defects, Strong
Effects, The Leading Edge, pp. 40-47, January 2009
Gutierrez, M., ino, L. E., and Nygrd, R. 2000. Stress-Dependent Permeability of a De-
Mineralised Fracture in Shale, Marine and Petroleum Geology 17, 895-907.
Hunt, J. M., Generation and Migration of Petroleum from Abnormally Pressured Fluid
Compartments, The American Association of Petroleum Geologists Bulletin. v.74, No 1, pp. 1-12
Hyman, L.A., Malek, D.J., Admire, C.A., Walls, J.D. 1991. The Effects of Microfractures on
Directional Permeability in Tight Gas Sands, SPE 21878, SPE Rocky Mountain Regional/Low
Permeability Reservoirs Symposium and Exhibition, April 15-17 1991, Denver, CO.
Javadpour, F., Fisher, D., and Unsworth, M. (2007). Nanoscale Gas Flow in Shale Gas Sediments,
Jour. Canadian Pet. Tech. Volume 46, No. 10, October, pp. 55-61.
Javadpour, F. (2009). Nanopores and Apparent Permeability of Gas Flow in Mudrocks (Shales
and Siltstone), Jour. Canadian Pet. Tech.Volume 48, No. 8, August, pp. 16-21.
Kazemi, H. 1969. Pressure Transient Analysis of Naturally Fractured Reservoirs with Uniform
Fracture Distribution, SPEJ 9 (4) 451-462. SPE 2516-A.
Kazemi, H., Merrill, Jr., L. S., Porterfield, L. L., and Zeman, P. R. 1976. Numerical Simulation of
Water-Oil Flow in Naturally Fractured Reservoirs, SPEJ 16 (6): 317-326. SPE 5719-PA.
Klinkenberg, L. J. (1941). The Permeability of Porous Media to Liquids and Gases, Drill. And
Prod. Prac. API. 200-213.
Larson, R. E., Higdon, J. J. L. 1986. Microscopic Flow Near the Surface of Two-Dimensional
Porous Media, Part 1. Axial Flow, Journal of Fluid Mechanics, vol. 166, pp. 449-472.
Liu C.Q., 1981. Exact Solution for the Compressible Flow Equations through a Medium with
Triple Porosity, Applied Mathematics and Mechanics 2 (4), pp. 457-462.
Liu C.Q., 1983. Exact Solution of unsteady Axisymmetrical Two-Dimensional Flow through
Triple Porosity Media, Applied Mathematics and Mechanics 4 (5), pp. 717-724.
Loucks, R. G., Reed, R. M., Ruppel, S. C., Jarvie, D. M. 2009. Morphology, Genesis, and
Distribution of Nanometer-Scale Pores in Siliceous Mudstones of the Mississippian Barnett Shale,
Journal of Sedimentary Research, 2009, vol. 79, pp. 848-861
Medeiros, F., Jr. Ozkan, E., and Kazemi, H. 2008. Productivity and Drainage Area of Fractured
Horizontal Wells in Tight Gas Reservoirs. SPE Reservoir Evaluation and Engineering, Vol. 11,
No. 5, October, pp. 902-911.
Min, K., Rutqvist, J., Tsang, C., and Lanru, J. 2004. Stress-dependent permeability of fractured
rock masses: a numerical study, Int J Rock Mech Min Sci (41): 1191-1210.
Mukherjee, H. and Economides, M. J. 1991. A Parametric Comparison of Horizontal and Vertical
Well Performance. SPE Formation Evaluation, June, pp. 209-216.
77

Najurieta, H. L. 1980. A Theory for Pressure Transient Analysis in Naturally Fractured
Reservoirs, Jour. Pet. Tech. July 241-1250.
Nelson, P., Pore-throat Sizes in Sandstones, Tight Sandstones, and Shales, AAPG Bulletin, v. 93,
no. 3, pp. 329340, March 2009
Olson, T. 2012. Its a Small World After All: Pores in Mudrocks, presented at the Denver
Geologist Study Group, January 2012
Ozkan, O., Raghavan, R. 1991. New Solutions for Well-Test-Analysis Problems: Part 1 -
Analytical Considerations, SPE 18615-PA.
Ozkan, O., Raghavan, R. 1991. Supplement to New Solutions for Well-Test-Analysis Problems:
Part 1--Analytical Considerations (Supplement to SPE 18615), SPE 23593-PA.
Ozkan, O., Raghavan, R. 1991. New Solutions for Well-Test-Analysis Problems: Part 2
Computational Considerations and Applications, SPE 18616-PA.
Ozkan, E., Brown, M., Raghavan, R., and Kazemi, H. (2009). Comparison of Fractured
Horizontal-Well Performance in Conventional and Unconventional Reservoirs, SPE 121290, SPE
Western Regional Meeting, 2426 March 2009, San Jose, California.
Pedrosa, Jr. 1986. Pressure Transient Response in Stress-Sensitive Formatoins, Paper SPE 15115
presented 56
th
California Regional Meeting of SPE, 2-4 April, Oakland, California.
Raghavan, R., and Ozkan, E. 2011. Flow in Composite Slabs, SPEJ 16 (2): 374-387, SPE
140748-PA.
Raghavan, R., and Chen, L. Y. 2004. Productivity Changes in Reservoirs With Stress-Dependent
Permeability, Paper SPE 88870 presented SPE Annual Technical Conference and Exhibition, 29
September-2 October, San Antonio, Texas.
Roy S., Raju, R., Chuang, H., Cruden, B., and Meyyappan, M., Modeling Gas Flow through
Microchannels and Nanopores, Journal of Applied Physics, v. 93, no. 8, pp. 4870-4878, April
2003
Serra, K., Reynolds, A. C., and Raghavan, R. 1983. New Pressure Transient Analysis Methods
for Naturally Fractured Reservoirs. Jour. Pet. Tech. (Dec.) 2271-2283.Stehfest, H. 1970.
Numerical Inversion of Laplace Transforms. Communications. ACM 13 (1). 47-49.
Slatt, R.M., OBrien N.R. 2011. Pore Types in the Barnett and Woodford Gas Shales:
Contribution to Understanding Gas Storage and Migration Pathways in Fine-grained Rocks,
AAPG Bulletin, V. 95 NO. 12, December 2011, pp. 2017-2030.
Soeder, D. J., Kappel, W. M. 2009. Water Resources and Natural Gas Production from the
Marcellus Shale, U.S. Geological Survey Fact Sheet, 2009-3032, pp. 6.
Van Everdingen, A., and Hurst, W. 1949. The Application of the Laplace Transformation to Flow
Problems in Reservoirs, SPE 949305-G.
Walsh J. B. (1981). Effect of Pore Pressure and Confining Pressure on Fracture Permeability. Int.
J. Rock Mech. Min. Sci. and Geomech. Abstr., 18, 3, 429-435.
78

Wang, J., Liu, Y. 2011. Simulation Based Well Performance Modeling in Haynesville Shale
Reservoir. SPE 142740, SPE Production and Operations Symposium, March 27-29 2011,
Oklahama City, Oklahama
Warren, J. E., and Root, P. J. 1963. The Behavior of Naturally Fractured Reservoirs. SPEJ 3 (3):
245 255, SPE 426-PA.

79

APPENDIX A
DERIVATION OF THE SOLUTION FOR FLOW IN A MICROFRACTURED
SPHERICAL MATRIX BLOCK
A composite spherical matrix block as shown in Fig. 4.2 is considered. The core of the
matrix block is homogeneous and the surface-layer of the matrix block has microfractures.
Consistent with the conventional dual-porosity formulations (Warren and Root, 1963, de Swaan,
1976, Kazemi, 1969), it is assumed that the pressure and flux are uniformly distributed on the
surface of each matrix block. To derive the flow solution for the composite matrix block, the flow
equations for the core and the surface layer of the matrix block are derived and the two solutions
at the boundary of the two regions are coupled.
Appendix A.1 Flow in the Matrix Core:
Mass balance for the flow of a real gas of density
g
! , viscosity
g
, and compressibility
g
c on a
spherical control volume of porosity
m
! in the matrix core yields
( ) ( )
2
2
1
; 0
g rm g m mc
r v r r
r r t
! ! "
# #
$ = % %
# #
. (A.1.1)
The radial component of the Darcy velocity in (A.1.1) is given by
m mc
rm c
g
k p
v
r
!

"
= #
"
. (A.1.2)
Substituting (A.1.2) into (A.1.1) and using the real gas equation of state,
mc g
g
g
p M
zR T
! = , (A.1.3)
one can obtain
2
2
1
; 0
m mc mc mc
c m mc
g
k p p p
r r r
r r z r t z
! "

# $ % &
% &
' ' '
( ) *+
= , , ) *
( ) *+ ) *
) *
- . ' ' '
( +
- . / 0
. (A.1.4)
Because
1 1 p p dz
p
z z p z dp
! "
! "
# $
% = & % # $
# $
# $
' (
' (
, (A.1.5)
80

1 1
g
dz
c
p z dp
= ! , (A.1.6)
1
f
c
p
!
!
"
=
"
, (A.1.7)
and
t g f
c c c = + , (A.1.8)
one can write (A.1.4) as follows:
2
2
1
; 0
m tm g
mc mc mc mc
mc
g c m g
c
p p p p
r r r
r r z r k z t
!
"
# $
% % %
& '
= ( (
& '
& '
% % %
) *
. (A.1.9)
Lets define a pseudopressure by
( )
0
2
p
g
p
m p dp
z
!
! =
"
. (A.1.10)
Then
2
2
1 1
; 0
mc mc
mc
m
m m
r r r
r r r t !
" #
$% $% $
= & & ' (
' (
) * $ $ $
(A.1.11)
where
( ) ( ) ( )
mc mc i i mc mc
m p m p m p ! = "
(A.1.12)
and
c m
m
m tm g
k
c
!
"
#
= . (A.1.13)
A solution of (A.1.11) subject to the following conditions can be obtained with the following:
The Initial condition is given by
( ) , , 0 0
mci m i
m r R t m ! = = ! = . (A.1.14)
The boundary condition at 0 r = is
81

( ) ( )
0
0, ,
mc m mc
m r R t m t finite ! = = ! = . (A.1.15)
The boundary condition at the interface between the core and the surface layer of the matrix block,
mc
r , is obtained from the continuity of the pressures given by
( ) ( ) , , , ,
mc mc mc m mf mf mc m
m p r r R t m p r r R t ! ! " # " # = = =
$ % $ %
, (A.1.16)
where ( ) , ,
mf mf m
m p r R t ! " #
$ %
and ( ) , ,
mf m
p r R t are the pseudopressure difference and the pressure,
respectively, in the fracture medium of surface layer of the matrix.
Using the dimensionless variables given by (4.1.2), (4.1.4), (4.1.5), and (4.1.6), (A.1.11) and
(A.1.14) through (A.1.16) can be written as
2
2
2 2
1 2
; 0
mcD mcD mcD mcD
D mD D mcD
D D D D D D D
m m m m
r r r
r r r r r r t
!
" #
$ $ $ $ $
% &
= + = ' '
% &
$ $ $ $ $
( )
, (A.1.17)
where
( )
fi m tm g f
mD
m f tf g m
i
c k
c k
! "
!
! "
= = , (A.1.18)
( ) , , 0 0
mcD D mD D
m r R t = = , (A.1.19)
( ) ( )
0
0, ,
mcD D mD D mcD D
m r R t m t finite = = = , (A.1.20)
and
( ) ( ) , , , ,
mcD D mcD mD D mfD D mcD mD D
m r r R t m r r R t = = = . (A.1.21)
To proceed with an analytical solution, it is assumed that
( )
( )
m tm g f
fi
i
mD mDi
mi f tf g m
i
c k
c k
!
"
" "
" !
# = = (A.1.22)
and linearize (A.1.17) as follows:
2
2
2 2
1 2 1
; 0
mD mD mD mD
D D mcD
D D D D D D mDi D
m m m m
r r r
r r r r r r t !
" #
$ $ $ $ $
% &
= + = ' '
% &
$ $ $ $ $
( )
. (A.1.23)
Lets consider the substitution
82

( ) ( ) , , , ,
mcD D mD D D mcD D mD D
w r R t r m r R t = . (A.1.24)
Then, (A.1.23) and (A.1.19) through (A.1.21) can be expressed in terms of
( ) , ,
mcD D mD D
w r R t ,
respectively, as follows:
2
2
1
mcD mcD
D mDi D
w w
r t !
" "
=
" "
, (A.1.25)
( ) , , 0 0
mcD D mD D
w r R t = = , (A.1.26)
( ) 0, , 0
mcD D mD D
w r R t = = , (A.1.27)
and
( ) ( ) ( ) , , , , , ,
mcD mcD mD D mcD mfD mcD mD D mfD mcD mD D
w r R t r m r R t w r R t = = . (A.1.28)
If the Laplace Transformation of (A.1.25), (A.1.27), and (A.1.28) are taken then,
2
2
0
mcD
mcD
D mDi
w s
w
r !
"
# =
"
, (A.1.29)
where (A.1.26) is used,
( ) 0, , 0
mcD D mD
w r R s = = (A.1.30)
and
( ) ( ) ( ) , , , , , ,
mcD mcD mD mcD mfD mcD mD mfD mcD mD
w r R s r m r R s w r R s = = . (A.1.31)
The solution of (A.1.29) is given by
( ) ( )
exp exp
mcD mDi D mDi D
w A s r B s r ! ! = " + . (A.1.32)
From (A.1.30), A B = ! . Then,
( ) ( ) ( )
exp exp sinh
mcD mDi D mDi D mDi D
w B s r s r C s r ! ! !
" #
= $ $ =
% &
. (A.1.33)
Using (A.1.31),
( )
( )
, ,
sinh
mfD mcD mD
mDi mcD
w r R s
C
s r !
= . (A.1.34)
83

Thus,
( )
( )
( )
( )
sinh
, , , ,
sinh
mDi D
mcD D mD mfD mcD mD
mDi mcD
s r
w r R s w r R s
s r
!
!
= . (A.1.35)
Finally, using (A.1.24), the solution given by (4.2.1) in the Section 4.2 is obtained.
Appendix A.2 Flow in the Matrix Surface Layer:
To model flow in the matrix surface layer, the transient dual-porosity idealization for a
naturally fractured porous medium suggested by de Swaan (1976) and Kazemi (1969) is used and
the surface layer by a linear system as shown in Figs. 4.3 and 4.4 is approximated.
Flow in the Surface-Layer Microfractures: For the geometry considered in Figs. 4.3 and
4.4, one-dimensional flow in the matrix fractures is described by
( ) ( ) ( )
, ;
g rmf g ms g mf mc m
v q r t r r r
r t
! ! ! "
# #
$ + = % %
# #
! , (A.2.1)
where
( ) ,
ms
q r t ! is the influx from matrix per unit volume of fracture in the matrix surface layer
per unit time. Following de Swaan O (1976), it was assumed that the flux from each surface of
the matrix slab is instantaneously and uniformly distributed in one-half the fracture slab adjacent
to the matrix surface. If
f mf mf
V A h =
!
denote the volume of each fracture slab, where
mf
A is the
area of the contact surface between the matrix and fracture, and
( ) 2,
ms mm
q h t ! = denote the
matrix efflux per unit time,
( ) ( )
( ) ( ) ( )
( ) 2,
, 2, 2 ,
, ,
2
mm
g m mm g
m ms
g m c
mf mf mf g
h t
r t q h t r t
k p
r t q r t
A h h
!
" ! "
" #
!
=
$ % & '
=
(
) * + ,
= = -
) * + ,
) *
(
+ ,
. / 0 1
! . (A.2.2)
Using (A.2.2) and
mf mf
rmf c
g
k p
v
r
!

"
= #
"
, (A.2.3)
and
mf g
g
g
p M
zR T
! = , (A.2.4)
84

(A.2.1) can be written as
( ) 2,
2
;
mm
mf mf m mf mf tmf g mf mf
ms
mc m
h t g mf mf g c mf g
p p k p c p p
p
r r r
r z r k h z k z t
!
"
! #
=
$ %
$ % & &
& &
' (
) = * * ' (
' ( ' (
' (
+ , & & & &
+ ,
, (A.2.5)
where (A.1.5) through (A.1.8) were also used. The initial condition is given by
( ) , 0
mf i
p r t p = = . (A.2.6)
One of the boundary conditions is the continuity of flux at the interface of the matrix core and the
matrix surface layer,
mc
r ; that is,
( )
( )
( )
( ) , ,
mc mc
g mf g mc
r t r t
q q ! ! = ! ! . Using the bulk fracture
permeability,
( )
mf mf mf mm mf mf mf mm
k k h h h k h h = + !
!
, (A.1.3), and (A.2.4), this condition is given
by
( ) ( ) , ,
mc mc
mf mf mf
mc mc
mf m
mm g g
r t r t
h p p
p p
k k
h z r z r
! " ! "
#
#
$ % $ %
=
$ % $ %
$ % $ %
# #
& ' & '
. (A.2.7)
The other condition is the continuity of pressure at the matrix-fracture interface at,
m
r , given by
( ) ( ) , ,
mf m f m
p r r t p R R t = = = . (A.2.8)
In terms of the pseudopressure defined by (A.1.10), (A.2.5) through (A.2.8) are written as follows,
respectively:
( ) 2,
2 1
;
mm
mf mf
m ms
mc m
h t mf mf mf
m m
k m
r r r
r r k h t
!
! "
=
# $ # $ %& %&
%& %
' = ( ( ) * ) *
) * ) *
+ , + , % % % %
, (A.2.9)
( ) , 0 0
mf
m r t ! = = , (A.2.10)
( ) ( ) , ,
mc mc
mf mf
mc
mf m
r t r t mm
h m
m
k k
h r r
! " ! " #$
#$
= % & % &
% & % &
' ( ' ( # #
, (A.2.11)
and
( ) ( ) , ,
mf m f m
m r r t m R R t ! = = ! = . (A.2.12)
In (A.2.9) through (A.2.12),
85

( ) ( ) ( )
mf mf i i mf mf
m p m p m p ! = " , (A.2.13)
( ) ( ) ( )
f f i i f f
m p m p m p ! = " , (A.2.14)
and
c mf
mf
mf tmf g
k
c
!
"
#
= . (A.2.15)
In terms of the dimensionless variables defined in (4.1.2) and (4.1.4),
( )
2
2
2 2
1,
4
;
D D
mfD mfD
m mm msD
mfD mcD D mD
D mf mf mm D D
t
m m
k h L m
r r r
r k h h t
!
"
!
=
# $
% %
%
& '
( = ) )
& '
% % %
* +
, (A.2.16)
where
( )
fi mf tmf g f
mfD
mf f tf g mf
i
c k
c k
! "
!
! "
= = , (A.2.17)
To linearize the equation, it is assumed that
( )
c mf
mf mfi
mf tmf g
i
k
c
!
" "
#
$ = , (A.2.18)
and define
fi
mfD mfDi
mfi
!
! !
!
" = . (A.2.19)
Then, (A.2.16), and (A.2.10) through (A.2.13) become, respectively,
( )
2
2
2 2
1,
4
;
D D
mfD mfD
m mm msD
mfDi mcD D mD
D mf mf mm D D
t
m m
k h L m
r r r
r k h h t
!
"
!
=
# $
% %
%
& '
( = ) )
& '
% % %
* +
. (A.2.20)
( ) , 0 0
mfD D D
m r t = = , (A.2.21)
( ) ( ) , ,
mcD D mcD D
mfD
m mm mcD
D mf mf D
r t r t
m
k h m
r k h r
! " ! "
#
#
$ % $ %
=
$ % $ %
# #
& ' & '
, (A.2.22)
and
86

( ) ( ) , ,
mfD mD D fD mD D
m r t m R t = . (A.2.23)
Flow in the Surface-Layer Matrix: Based on the representation of the surface-layer as
slabs of matrix and fracture media (Fig. 4.4), one-dimensional flow in the surface-layer matrix is
defined by
( ) ( )
; 0 2
g m g m mm
v h
t
!
" " # !
!
$ $
% = & &
$ $
. (A.2.24)
Substituting
m ms
m c
g
k p
v
!
"
!
#
= $
#
, (A.2.25)
and
ms g
g
g
p M
zR T
! = , (A.2.26)
one can write (A.2.24) as
; 0 2
m tm g
ms ms ms ms
mm
g c m g
c
p p p p
h
z k z t
!
"
" " #
$ %
& & &
' (
= ) )
' (
' (
& & &
* +
, (A.2.27)
where (A.1.5) through (A.1.8) were used. The initial and boundary conditions are given by
( ) , 0
ms i
p t p ! = = , (A.2.28)
( ) 0,
0
ms
t
p
!
!
=
"
=
"
, (A.2.29)
( ) ( ) 2, ,
ms mm mf
p h t p r t ! = = , (A.2.30)
In terms of the pseudopressure defined by (A.1.10), (A.2.27) through (A.2.30) are written as
follows, respectively:
1
; 0 2
ms ms
mm
m
m m
h
t
!
! ! "
# $
%& %& %
= ' ' ( )
( )
* + % % %
, (A.2.31)
( ) , 0 0
ms
m t ! " = = , (A.2.32)
87

( ) 0,
0
ms
t
m
!
!
=
"#
=
"
, (A.2.33)
and
( ) ( ) 2, ,
ms mm mf
m h t m r t ! " = = " , (A.2.34)
where
( ) ( ) ( )
ms ms i i ms ms
m p m p m p ! = " . (A.2.35)
Defining
2
D
mm
h
!
! = , (A.2.36)
and using the dimensionless variables defined in (4.1.2) and (4.1.4), following is obtained
2
2
2 2
; 0 1
4
mm fi
msD msD
D
D m D
h
m m
L t
!
"
" !
# #
= $ $
# #
, (A.2.37)
( ) , 0 0
msD D D
m t ! = = , (A.2.38)
( ) 0,
0
D D
msD
D
t
m
!
!
=
"
=
"
, (A.2.39)
and
( ) ( ) 1, ,
msD D D mfD D D
m t m r t ! = = . (A.2.40)
To proceed with an analytical solution, it is assumed that
( )
c m
m mi
m tm g
i
k
c
!
" "
#
$ = , (A.2.41)
and defined
( )
c mf
mfi
mf tmf g
i
k
c
!
"
#
= , (A.2.42)
and
88

( )
( )
f tf g mf
mfi
i
mfDi
fi mf tmf g f
i
c k
c k
!
"
"
" !
= = . (A.2.43)
Following Serra et al. (1983), it is also defined that
2
2
12
m mm
m
mm mf mf
k h L
h k h
!
" #
$ %
=
$ %
$ %
& '
(A.2.44)
and
( )
( )
t mm
mi
m
t mf
mfi
c h
c h
!
"
!
= . (A.2.45)
Then, (A.2.37) becomes
2
2
3
; 0 1
msD m msD
fmDi D
D m D
m m
t
!
" #
# $
% %
= & &
% %
. (A.2.46)
Taking the Laplace transformation of (A.2.46), (A.2.39), and (A.2.40), following is obtained,
respectively,
2
2
3
=0
msD m
mfDi msD
D m
m
s m
!
"
# $
%
&
%
, (A.2.47)
where (A.2.38) is used,

( ) 0,
0
D
msD
D
s
m
!
!
=
"
=
"
, (A.2.48)
and
( ) 1,
msD D mfD
m s m ! = = . (A.2.49)
The general solution of (A.2.47) is given by
3 3
exp exp
m m
msD mfDi D mfDi D
m m
m A s B s
! !
" # " #
$ $
% & % &
' ( ' (
= ) +
' ( ' (
* + * +
. (A.2.50)
Using (A.2.48) and (A.2.49), the solution given in the text by (4.2.3) is obtained.
89

Coupled Flow Solution for the Matrix Surface Layer: Using (A.2.44) and taking the
Laplace transformation of (A.2.20), following is obtained

( )
2
2
1,
0
3
D D
mfD
m msD
mfDi mfD
D D
t
m
m
sm
r
!
"
#
!
=
$ %
&
&
' (
) ) =
' (
& &
* +
, (A.2.51)
where (A.2.21) is used. Using (4.2.3), one can write (A.2.51) as follows

2
2
0
mfD
m mfD
D
m
u m
r
!
" =
!
. (A.2.52)
In (A.2.52), following is defined

( )
m m
u sf s = , (A.2.53)
where

( )
3
1 tanh
3
m m m
m mfDi mfDi
mfDi m
f s s
s
! " "
# #
# !
$ %
& '
( ) * +
= +
( ) * +
( ) , -
. /
. (A.2.54)
Taking the Laplace transforms of (A.2.22), following is obtained

( )
( )
,
,
mcD
mfD
mf mfD mcD
D
r s
m
f m r s
r
! "
#
$ %
=
$ %
#
& '
, (A.2.55)
where following is used
( )
( )
( )
,
1
coth ,
mcD
mcD
mDi mDi mcD mfD mcD
D mcD
r s
m
s s r m r s
r r
! !
" # $ %
&
' ( ) *
= +
' ( ) *
&
, - . /
, (A.2.56)
and
( )
2
2
coth 1
12
mm m
mf mDi mcD mDi mcD
mcD
h
f s r s r
r L
!
" "
# $
= %
& '
. (A.2.57)
90

Also taking the Laplace transform of (A.2.23), one would have
( ) ( ) , ,
mfD mD fD mD
m r s m R s = . (A.2.58)
The general solution of (A.2.52) is given by
( ) ( )
exp exp
mfD m D m D
m A u r B u r = ! + . (A.2.59)
From (A.2.55) and (A.2.59), one would have
( )
( ) ( )
,
exp exp
mcD
mfD
mf m mcD m mcD
D
r s
m
f A u r B u r
r
! "
#
$ % & '
= ( +
) *
$ %
#
+ ,
, (A.2.60)
which yields
( )
( )
( )
exp 2
m mf
m mcD
m mf
u f
B A u r
u f
+
= !
!
. (A.2.61)
Substituting (A.2.61) into (A.2.59),
( )
( )
( )
( )
1 exp 2 exp
m mf
mfD m D mcD m D
m mf
u f
m u r r A u r
u f
! "
+
# #
$ %
= + & &
' (
) *
# #
&
+ ,
. (A.2.62)
Using (A.2.58) and (A.2.62),
( )
( )
( )
( )
( )
exp ,
1 exp 2
m mD fD mD
m mf
m mD mcD
m mf
u r m R s
A
u f
u r r
u f
=
! "
+
# #
$ %
+ &
' (
) *
# # &
+ ,
. (A.2.63)
Substituting (A.2.63) into (A.2.62), the particular solution given below is obtained (also (4.2.4) in
text):
( )
( ) ( )
( )
( ) ( )
( )
( ) ( )
exp 2
, exp ,
exp 2
m mf m mf m D mcD
mfD D m mD D fD mD
m mf m mf m mD mcD
u f u f u r r
m r s u r r m R s
u f u f u r r
! "
# + + #
$ %
! "
= #
$ %
! "
# + + #
$ %
.
(A.2.64)

91

APPENDIX B
DERIVATION OF THE SOLUTION FOR RADIAL FLOW IN THE
MACROFRACTURE NETWORK
Consider a cylindrical, dual-porosity reservoir. Mass balance for flow of a real gas of
density
g
! , viscosity
g
, and compressibility
g
c on a fracture control volume of porosity
f
!
yields
( ) ( ) ( )
1
,
g Rf g mf g f
R v q R t
R R t
! ! ! "
# #
$ + =
# #
! , (B.1.1)
where
( ) ,
mf
q R t ! is the influx from matrix per unit volume of fracture per unit time. Note that it is
assumed that fluid transfer from matrix to fractures is through the matrix fractures only; that is,
when flux from the matrix is considered, only the flux from matrix fractures is considered.
Following de Swaan O (1976), it is assumed that the flux from each matrix block is
instantaneously and uniformly distributed in one-half the fracture volume that envelopes the
matrix block,
f
V
!
. If we let
( ) , ,
mf m m
q r r R t = denote the matrix outflux per unit time,
( ) ( )
( ) ( )
( )
( )
( )
2
, , , ,
, ,
4
, ,
2 2
m m m m
g mf g mf
g mf c mf
f f g
r r R t r r R t
R t q R t p
R t q R t r k
V V r
! !
"
! #

= =
$ % & '
(
) * + ,
= = -
) * + ,
) *
(
+ ,
. / 0 1
!
!
! !
(B.1.2)
where
mf
k
!
is the bulk permeability of the fracture matrix defined by (see Fig. 4.4)
mf
mf mf
mm
h
k k
h
=
!
. (B.1.3)
If it is assumed that the fracture volume that envelopes the matrix block,
f
V
!
, has a uniform
average thickness of
f
h , and that
f
h is small compared with the dimensions of the matrix block,
m
r , one can approximate
2
1
4
2 2 2
f f
f m m
h h
V A r ! " =
!
. (B.1.4)
Using (B.1.3) and (B.1.4), (B.1.2) becomes
92

( ) ( )
( ) , ,
2
, ,
m m
mf g mf
g mf c mf
f mm g
r r R t
h p
R t q R t k
h h r
!
! "

=
# $
%
& '
= (
& '
& '
%
) *
! . (B.1.5)
The radial component of the velocity vector for flow in fractures is given by
f f
Rf c
g
k p
v
R
!

"
= #
"
. (B.1.6)
Substituting (B.1.5), (B.1.6) and the real gas equation of state,
f g
g
g
p M
zR T
! = , (B.1.6a)
into (B.1.1), following is obtained
( ) , ,
2
1
m m
f f mf mf f mf f tf f f
g f f mm g c f
r r R t
p p k h p p c p p
R
R R z R k h h z r k z t
!
"
=
# $ # $
% % %
%
& ' & '
( =
& ' & '
& ' & '
% % % %
) * ) *
, (B.1.7)
where (A.1.5) through (A.1.8) are also used.
Assume that initially
( ) , 0
f i
p R t p = = . (B.1.8)
For a vertical, fully penetrating well of radius
w
R located at the center of the reservoir, 0 R = , the
inner boundary condition is
2 2
w w
g g f f f ft sc f f
sc ft c c
gsc gsc g fsc g
R R R R
k p k h T p p
q q h R R
R p T z R
! !
" # "#
! !
= =
$ % $ % & &
= = = ' ( ' (
' ( ' (
& &
) * ) *
, (B.1.9)
where production at a constant surface rate of,
sc
q , is considered, the effects of wellbore storage
is ignored, and no flow from matrix to wellbore as per the dual-porosity assumption is assumed.
In (B.1.9),
ft
h is the total thickness of the wellbore and fracture intersections.
For the outer boundary condition, one can use an infinite-acting system. Then, the outer boundary
condition is given by
( ) ,
f i
p R t p !" = . (B.1.10)
93

In terms of the pseudopressure defined by (A.1.10), (B.1.7) through (B.1.10) are given,
respectively, by
( ) , ,
2
1 1
m m
f mf mf mf f
r r R t f f mm f
m k h m m
R
R R R k h h r t !
=
" # " # $% $% $%
$
& = ' ( ' (
' ( ' (
) * ) * $ $ $ $
, (B.1.11)
( ) , 0 0
f
m R t ! = = , (B.1.12)
w
fi ft sc f
sc
fsc
R R
k h T m
q R
p T R
!
=
"# $ %
= &
' (
' (
"
) *
, (B.1.13)
and
( ) , 0
f
m R t ! "# = . (B.1.14)
In (B.1.11),
c f
f
f tf g
k
c
!
"
#
= . (B.1.15)
In terms of the dimensionless variables, (4.1.2) and (4.1.4) through (4.1.6), one can write (B.1.11)
through (B.1.14) as follows, respectively:
( ) , ,
2
1
D mD mD D
fD mf mf mfD fD
D
D D D f f mm D D
r r R t
m k h L m m
R
R R R k h h r t
=
! " ! "
# # #
#
$ % $ %
& =
$ % $ %
# # # #
' ( ' (
, (B.1.16)
( ) , 0 0
fD D D
m R t = = , (B.1.17)
1
D wD
fD
D
D
R R
m
R
R
=
! " #
$ =
% &
% &
!
' (
, (B.1.18)
and
( ) , 0
fD D D
m R t !" = . (B.1.19)
To linearize (B.1.16), it is also assumed
( )
1
fi m tm g f
fD
f f tf g f
i
c k
c k
! "
!
! "
= = # . (B.1.20)
94

Taking the Laplace transform of (B.1.16), (B.1.18) and (B.1.19), respectively,
( ) , ,
2
1
0
mD mD
fD mf mf mfD
D fD
D D D f f mm D
r R s
m k h L m
R sm
R R R k h h r
! " ! "
# #
#
$ % $ %
& & =
$ % $ %
# # #
' ( ' (
, (B.1.21)
where (B.1.17) is used,
1
D wD
fD
D
D
R R
m
R
R s
=
! " #
= $
% &
% &
!
' (
, (B.1.22)
and
( ) , 0
fD D
m R s !" = . (B.1.23)
From (4.2.4),
( )
( )
, ,
5
,
mD mD
mfD
mm
f m fD mD
D mf mD
r R s
m
h
f f sm R s
r h r
! "
#
$ %
= &
$ %
#
' (
, (B.1.24)
where
( ) ( )
( )
( ) ( )
( )
exp 2
5 exp 2
m mf m mf m mD mcD
mf mD
f
m mm m mf m mf m mD mcD
u f u f u r r
h r
f
u h u f u f u r r
! "
# $
% &
' ' + '
( ( ) *
+ ,
=
- .
+ ,
( ( % &
' + + '
/ 0
) * 1 2
. (B.1.25)
Also given as (4.2.14) in the text. (B.1.21) can be written as follows:
2
2
10
1
1 0
fD mf m
D f m fD
D D D f f m
m k r L
R f f sm
R R R k h r
! "
! "
#
#
$ %
$ %
+ & =
$ %
$ %
$ %
# #
' (
' (
. (B.1.26)
Following the formulation of the standard dual-porosity models (Warren and Root, 1963, de
Swaan, 1976, Kazemi, 1969), lets define
( )
( )
( )
3 2
2 2 2 2
2 2 2
4 3 2 2
15
10
2 4 2 3 3
mf m mf m mf m mf m mf m
f f f m f f f m f f f f m
k V k r k r k r k r L
L L L L
k V k r h k h r k h k h r
!
" # # #
!
$ % $ % & ' & '
( ) ( ) * + * +
= = = = =
( ) ( ) * + * +
* + * +
( ) ( )
, - , - . / . /
.
(B.1.27)
95

In (B.1.27), the assumption that the flux from each matrix block is instantaneously and uniformly
distributed in one-half the fracture volume that envelopes the matrix block,
f
V
!
, (de Swaan, 1976)
is used and the shape factor is defined as, !
( )
( )
2
2
2 15
m m m
n
n V A r
!
+
= = . (B.1.28)
In the second equality of (B.1.29), the number of orthogonal fractures sets 3 n = is used for
spherical matrix blocks. Then from (B.1.26) and (B.1.27),
1
0
fD
D fD
D D D
m
R um
R R R
! "
#
#
$ %
& =
$ %
# #
' (
, (B.1.29)
where
( ) u sf s = , (B.1.30)
also (4.2.12) in the text and
( ) 1
f m
f s f f ! = " . (B.1.31)
The general solution of (B.1.29) is given by
( ) ( ) 0 0 fD D D
m AK uR BI uR = + , (B.1.32)
and, from (B.1.22) and (B.1.23), the particular solution is obtained as
( )
( )
0
1
D
fD
wD wD
K uR
m
s uR K uR
= . (B.1.33)
which is also given by (4.2.11) in the text. Note that, in the above derivations, one can substitute
the following relations:
( )
( )
2 2
12 15 fi t f
mi m
mDi
mi t m mD m mmD
fi
c k
c k r h
! "
# #
!
! " $ $
= = = (B.1.34)
and
96

2
15 fi
mi
mfDi mDi
mfi mfi mD
r
!
! "
! !
! ! #
= = = . (B.1.35)
where
( )
( ) ( )
( ) ( )
( ) ( )
( )
( )
3
2
4 3 2
2 4 2 3
t m t m t m
mfi mfi mfi
t f t m f t f
fi fi fi
c V c r c r
c V c r h c h
! ! " !
#
! ! " !
= = = (B.1.36)

97

APPENDIX C
DERIVATION OF DUAL-MECHANISM DUAL-POROSITY MODEL FOR SHALE-GAS
RESERVOIRS
Here the derivation of the dual-mechanism dual-porosity for a radial flow system in the
reservoir is demonstrated. A spherical geometry for the matrix blocks is considered. The
schematic of the dual-porosity system considered here is shown in Fig. 4.9.
Appendix C.1 Dual-Mechanism Flow in a Spherical Matrix Block
Consider flow in a spherical matrix block of radius,
m
r . Assume that the pressure and
flux are uniformly distributed on the surface of the matrix block. Mass balance for the flow of a
real gas of density
g
! , viscosity
g
, and compressibility
g
c on a spherical control volume of
porosity
m
! yields
( ) ( )
2
2
1
g rm g m
r v
r r t
! ! "
# #
$ =
# #
. (C.1.1)
Using the approach taken by Ertekin et al. (1986),the radial component of the matrix flow
velocity,
rm
v , in (C.1.1) is defined as the sum of the radial components of the Darcy velocity,
prm
v , and the slip velocity,
srm
v ; that is,
rm prm srm
v v v = + . (C.1.2)
The radial component of the Darcy velocity in (C.1.2) is given by
m m
prm c
g
k p
v
r
!

"
= #
"
. (C.1.3)
The radial component of the slip velocity is given by the modified form of the Ficks Law
(Ertekin et al., 1986) as follows:
g m
m
srm c
g
M D
C
v
r
!
"
#
= $
#
. (C.1.4)
In (C.1.4)
g
M is the molecular weight of the gas,
m
D is the diffusivity coefficient for the matrix,
and
m
C is the concentration. Then,
98

g m
m m m
rm c
g g
M D
k p C
v
r r
!
"
# #
= $ $
# #
. (C.1.5)
Substituting (C.1.5) into (C.1.1) and using the real gas equation of state,
g
g
g
pM
zR T
! = , (C.1.6)
we obtain
2
2
1
m m m m m
c g m m
g
k p p C p
r R TD
r r z r r t z
! "

# $ % &
% &
' ' ' '
( ) *+
+ = ) *
( ) *+ ) *
) *
, - ' ' ' '
( +
, - . /
. (C.1.7)
From the real gas law, molar concentration for single phase flow is
g
m
m
g g
p
C
M zR T
!
= = . (C.1.8)
Then, (C.1.7) becomes
2
2
1 1
m m m m m m m m m
c m
g m m m m
k p p p p p p z
r D
r r z r r z z p p z p t
! !
"
!
# $ % &
% &
' ( ' (
) ) ) ) ) ) * *
+ ,
+ ,
+ = + - . - .
/ 0
- . - . + ,
+ ,
* * 1 2 1 2 ) ) ) ) ) )
+ , 3 4
3 4 5 6
. (C.1.9)
Because
1 1 p p dz
p
z z p z dp
! "
! "
# $
% = & % # $
# $
# $
' (
' (
, (C.1.10)
1 1
g
dz
c
p z dp
= ! , (C.1.11)
1
f
c
p
!
!
"
=
"
, (C.1.12)
and
t g f
c c c = + , (C.1.13)
one can write (C.1.9) as follows:
99

2
2
1
m m m m m
c m g m tm
g
k p p p p
r D c c
r r z r z t
! "

# $ % &
' ' '
( ) * +
+ =
( ) * +
) *
' ' '
( +
, - . /
. (C.1.14)
Rearranging (C.1.14), we obtain
2
2
1
1
m g g
m m m m m
c m tm
g m
D c
k p p p p
r c
r r k z r z t

! "

# $
% &
' ' '
( ) * +
+ =
( ) * +
' ' '
( ) , -
. /
. (C.1.15)
Following Ertekin et al. (1986), we define an apparent gas slippage term by
m g g m
am
m
D c p
b
k

= , (C.1.16)
and write
2
2
1
1
m am m m m m
c m tm
g m
k b p p p p
r c
r r p z r z t
! "

# $
% &
' ' '
( ) * +
+ =
( ) * +
' ' '
( ) , -
. /
. (C.1.17)
Note that (C.1.17) includes an apparent permeability,
1
am
am m
m
b
k k
p
! "
# $
= +
# $
% &
, (C.1.18)
in the form defined by Klinkenberg (1941). As indicated by (C.1.16), however,
am
b is not a
constant; it is a function of pressure.
Lets define a pseudopressure by
( )
0
2 1
m
p
am
m m
g
b p
m p dp
p z
! "
#
$ %
# = +
$ %
#
& '
(
. (C.1.19)
Then, we have
( )
0
2 1 2 1
m
p
m m am m am m m
m g m g
m p b p b p p p
dp
r p p z r p z r
! "
# $ # $
% & % % %
' ( ) * ) *
& = + = +
' ( ) * ) *
& % % % %
+ , ' ( + ,
- .
/
, (C.1.20)
( )
0
2 1 2 1
f
p
f f f f f
am am
f g f g
m p p p p
b b p
dp
t p p z t p z t
! " # $
# $
% % %
& %
' ( ) *
' (
& = + = +
' ( ) *
' (
' (
& % % % %
+ , ) *
- . + ,
/
, (C.1.21)
100

and, substituting into (C.1.17),
2
2
1 1
m m
m
m m
r
r r r t !
" #
$% $% $
= & '
& '
( ) $ $ $
(C.1.22)
where
( ) ( ) ( )
m m mi i m m
m p m p m p ! = " (C.1.23)
and
1
am
c m
m c am
m
m tm g m tm g
b
k
p k
c c
!
!
"
# #
$ %
& '
+
& '
( )
= = . (C.1.24)
The solution of (C.1.22) subject to the following conditions is the objective:
The initial condition is given by
( ) , , 0 0
m m
m r R t ! = = . (C.1.25)
The boundary condition at 0 r = is
( ) ( )
0
0, ,
m m m
m r R t m t finite ! = = ! = . (C.1.26)
The boundary condition at the surface of the matrix block,
m
r , is obtained from the continuity of
the matrix and fracture pressures given by
( ) ( ) , , ,
m m m f m
p r r R t p R t = = ! , (C.1.27)
where ( ) ,
f m
p R t ! is the surface-averaged fracture pressure at
m
r r = (the surface of the matrix
block); that is,
( ) ( )
( )
( )
2
2
2
0 0
0
1
, ,
1
cos , sin
4
1
cos , sin
2
f m f m
S
m
f m m
m
f m
p R t p R t dA
A
p R t r d d
r
p R t d
! !
!
" " # "
!
" " "
=
= +
= +
$
$ $
$
!
, (C.1.28)
101

where
m
A is the surface area of a matrix block and
m
dA is the differential element of the matrix
surface. The assumption of uniform pressure distribution on the surface of the matrix block is
based on the premise that the radius of investigation is larger than the dimension of the matrix
blocks. Therefore, the time for this assumption to be valid increases as the size,
m
r , of the matrix
blocks increases. Equation (C.1.27) can be expressed in terms of pseudopressure as follows:
( )
( )
( ) ( ) ( )
( )
,
, , ,
,
, , 2 1
2 1 2 2
2 2 1 1
i
f m
i i i
f f
f m f m f m
i
f f f
f m
p
am
m m m
p R t
g
p p p
d p d p
am
p R t p R t p R t
g g g
p
d p d p d
am
p R t
g
b p
m r r R t dp
p z
b p p p
dp e dp e dp
p z z z
b p
e dp e e
z p

! ! " # " #
! ! " # # "
$ %
!
& '
! # = = +
& '
!
( )
$ %
! ! !
& '
! ! ! = + " +
& '
!
( )
* + $ %
!
, & '-
! = " " +
, & '-
!
( ) . /
0
0 0 0
0
!
! ! !
! ( ) ,
i
f m
p
p
p R t
g
p
dp
z
! #
!
!
0
!
.

(C.1.29)
Rearranging (C.1.29) to obtain
( )
( ) ( ) ( )
( ) ( )
( )
, , ,
, ,
, , 2 1 2 2
2 2 1 1
,
i i i
f f
f m f m f m
i i
f f f
f m f m
p p p
d p d p
am
m m m
p R t p R t p R t
g g g
p p
d p d p d p
am
p R t p R t
g g
f m b
b p p p
m r r R t dp e dp e dp
p z z z
b p p
e dp e e dp
z p z
m R t m


! ! " # " #
! ! ! " # # " #
$ %
! ! !
& '
! ! ! # = = + " +
& '
!
( )
* + $ %
! !
, & '-
! ! = " " +
, & '-
!
( ) . /
= # " #
0 0 0
0 0
! ! !
! !
! !
( ) ,
m
R t
,

(C.1.30)
where
( )
( ) ,
, 2
i
f
f m
p
d p
f f m
p R t
g
p
m p R t e dp
z
! " #
!
! $ % # # =
& '
(
!
! ! , (C.1.31)
and
( )
( )
( ) ( )
,
, 2 1 1
, , ,
i
f f
f m
p
d p d p
am
b m
p R t
g
f m m m m
b p
m R t e e dp
p z
m R t m r r R t

! ! " # "
$ % & '
!
( ) *+
! " = # +
( ) *+
!
, - . /
= " # " =
0
!
!
!
. (C.1.32)
In (C.1.30) through (C.1.32),
f
d is the constant used in the definition of a stress-dependent
natural fracture permeability by (Raghavan and Chin, 2004)
102

( )
exp
f f
d p
f fi f fi f fi
k k d p p k e
! "
# $ = ! ! =
% &

(C.1.33)
and
p
f is the mean value of ( )
p
f p which should be defined for the pressure interval between
f
p and
i
p .
Using the dimensionless variables given in (4.1.2) and (4.1.4) through (4.1.6), (C.1.22), (C.1.25),
(C.1.26), and (C.1.29) can be written as
2
2
2 2
1 2
mD mD mD mD
D mD
D D D D D D D
m m m m
r
r r r r r r t
!
" #
$ $ $ $ $
% &
= + =
% &
$ $ $ $ $
' (
, (C.1.34)
where
( )
( )
1
fi m tm g fi m tm g fi
mD
m f tf g am am
i
f tf g m
i
m
c k c k
c k b
c k
p
! " "
!
! "
"
= = =
# $
% &
+
% &
' (
, (C.1.35)
( ) , , 0 0
mD D mD D
m r R t = = , (C.1.36)
( ) ( )
0
0, ,
mD D mD D mD D
m r R t m t finite = = = , (C.1.37)
and
( ) ( ) ( ) 1, , , ,
mD D mD D fD mD D bD mD D
m r R t m R t m R t = = ! ! ! . (C.1.38)
To proceed with an analytical solution, we assume that
( )
( )
( )
( )
1
m tm g fi m tm g fi
fi
i i
mD mDi
mi f tf g ami ami
i
f tf g m
i
i
c k c k
c k b
c k
p
! !
"
" "
" !
!
# = = =
$ %
& '
+
& '
( )
, (C.1.39)
and linearize (C.1.34) as follows:
2
2
2 2
1 2
mD mD mD mD
D mDi
D D D D D D D
m m m m
r
r r r r r r t
!
" #
$ $ $ $ $
% &
= + =
% &
$ $ $ $ $
' (
. (C.1.40)
given as (4.3.16) in text. Lets consider the substitution
( ) ( ) , , , ,
mD D mD D D mD D mD D
u r R t r m r R t = . (C.1.41)
103

From (C.1.41), one can write
mD mD
mD D
D D
u m
m r
r r
! !
= +
! !
, (C.1.42)
2 2
2 2
1 2
mD mD mD
D D D D D
u m m
r r r r r
! ! !
= +
! ! !
, (C.1.43)
and
1
mD mD
D D D
u m
r t t
! !
=
! !
. (C.1.44)
Thus, (C.1.40) can be expressed in terms of
( ) , ,
mD D mD D
u r R t as follows:
2
2
mD mD
mDi
D D
u u
r t
!
" "
=
" "
. (C.1.45)
The Initial condition, (C.1.36), becomes
( ) , , 0 0
mD D mD D
u r R t = = , (C.1.46)
and the boundary conditions, (C.1.37) and (C.1.38), become, respectively,
( ) 0, , 0
mD D mD D
u r R t = = (C.1.47)
and
( ) ( ) ( ) , , , ,
mD D mD mD D fD mD D bD mD D
u r r R t m R t m R t = = ! ! ! . (C.1.48)
If the Laplace Transformation of (C.1.45), (C.1.47), and (C.1.48) is taken, following is obtained,
respectively,
2
2
0
mD
mDi mD
D
u
s u
r
!
"
# =
"
, (C.1.49)
where (C.1.46)is used,
( ) 0, , 0
mD D mD
u r R s = = (C.1.50)
and
104

( ) ( ) ( ) , , , ,
mD D mD mD fD mD bD mD
u r r R s m R s m R s = = ! ! ! . (C.1.51)
The solution of (C.1.49) is given by
( ) ( )
exp exp
mD mDi D mDi D
u A s r B s r ! ! = " + . (C.1.52)
From (C.1.50), A B = ! . Then,
( ) ( ) ( )
exp exp sinh
mD mDi D mDi D mDi D
u B s r s r C s r ! ! !
" #
= $ $ =
% &
. (C1.53)
Using (C.1.51),
( ) ( )
( )
, ,
sinh
fD mD bD mD
mDi mD
m R s m R s
C
s r !
"
=
! !
. (C.1.54)
Thus,
( )
( )
( )
( ) ( )
sinh
, , , ,
sinh
mDi D
mD D mD fD mD bD mD
mDi mD
s r
u r R s m R s m R s
s r
!
!
" # = $
% &
! ! . (C.1.55)
Finally, using (C.1.41),
( )
( )
( )
( ) ( )
sinh
, , , ,
sinh
mD mDi D
mD D mD fD mD bD mD
D mDi mD
r s r
m r R s m R s m R s
r s r
!
!
" # = $
% &
! ! . (C.1.56)
If following is defined,
( )
( )
( )
,
,
,
bD mD
p mD
fD mD
m R s
f R s
m R s
=
!
!

(C.1.57)
then,
( )
( )
( )
( ) ( )
sinh
, , 1 , ,
sinh
mD mDi D
mD D mD p mD fD mD
D mDi mD
r s r
m r R s f R s m R s
r s r
!
!
" # = $
% &
!

(C.1.58)

given as (4.3.18) in the text.
105

Appendix C.2 Flow in Fracture Network with Stress-Dependent Permeability
Consider a cylindrical, dual-porosity reservoir. Mass balance for flow of a real gas of density
g
! ,
viscosity
g
, and compressibility
g
c on a fracture control volume of porosity
f
! yields
( ) ( ) ( )
1
,
g Rf g m g f
R v q R t
R R t
! ! ! "
# #
$ + =
# #
! , (C.2.1)
where
( ) ,
m
q R t ! is the influx from matrix per unit volume of fracture per unit time. Following de
Swaan O (1976), it is assumed that the flux from each matrix block is instantaneously and
uniformly distributed in one-half the fracture volume that envelopes the matrix block,
f
V
!
. If we
let
( ) , ,
m m m
q r r R t = denote the matrix outflux per unit time, then
( ) ( )
( ) ( )
( )
( )
( )
2
, , , ,
, ,
4
, ,
2 2
m m m m
g m g
m
g m c am
f f g
r r R t r r R t
R t q R t
p
R t q R t r k
V V r
! !
"
! #

= =
$ % & '
(
) * + ,
= = -
) * + ,
) *
(
+ ,
. / 0 1
!
! !
, (C.2.2)
where (C.1.5) is used to define the radial component of the velocity vector in matrix. If it is
further assumed that the fracture volume that envelopes the matrix block,
f
V
!
, has a uniform
average thickness of
f
h , and that
f
h is small compared with the dimensions of the matrix block,
m
r , it can be approximated
2
1
4
2 2 2
f f
f m m
h h
V A r ! " =
!
. (C.2.3)
Using (C.2.3), (C.2.2) becomes
( ) ( )
( ) , ,
2
, ,
m m
g
m
g m c am
f g
r r R t
p
R t q R t k
h r
!
! "

=
# $
%
& '
= (
& '
& '
%
) *
! . (C.2.4)
(C.2.4) will be used to couple matrix and fracture flow equations later.
The radial component of the velocity vector for flow in fractures is given by
( )
exp
fi f
Rf pRf f fi f
g
k p
v v d p p
R
!
" # = = $ $ $
% &
!
. (C.2.5)
106

In (C.2.5) a stress-dependent permeability is considered and, following Raghavan and Chin
(2004), the fracture permeability is defined by
( )
exp
f f
d p
f fi f fi f fi
k k d p p k e
! "
# $ = ! ! =
% &
. (C.2.6)
The expression in (C.2.5) corresponds to Type I rocks (Raghavan and Chin, 2004). Other
relationships defining the stress-dependent permeability could also be used in the derivations
(Walsh, 1981, Gutierrez et al., 2000, Raghavan and Chin, 2004, Chipperfield et al., 2007).
Substituting (C.2.5) into (C.2.1) and using the real gas equation of state, (C.1.6),
1
1
f f
d p
fi f f f f
c m f
g
f f f f f
f f f f
k e p p p p
R q
R R z R z t z
p p p
z
z p p z p t
! "

" "
"
# $ % & ' (
' ( )
) )
* + , -
+ = + ,
* + , - + ,
+ ,
. / ) ) )
* -
. / 0 1
% &
' ( ) )
)
* -
= + + ,
+ , * -
. / ) ) )
* -
0 1
!
.
(C.2.7)
Also using (C.1.10) through (C.1.13), (C.2.7) can be written as follows:
1
f f
d p
f f f f tf f f
m
g fi fi
p p p c p p
e
R q
R R z r k z k z t
!

" #
$ % & '
( (
(
) * + ,
+ =
) * + ,
* +
( ( (
) ,
- . / 0
! . (C.2.8)
Assume that initially
( ) , 0
f fi i
p R t p p = = = . (C.2.9)
For a vertical, fully penetrating well of radius
w
R located at the center of the reservoir, 0 R = , the
inner boundary condition is
2 2
f f
w w
d p g g f f fi ft sc f f
sc ft c c
gsc gsc g fsc g
R R R R
k p k h T p p
q q h R Re
R p T z R
! !
" # "#
! !
$ %
= =
& ' & ' ( (
= = = ) * ) *
) * ) *
( (
+ , + ,
, (C.2.10)
where production at a constant surface rate of,
sc
q , is considered, the effects of wellbore storage
is ignored, and no flow from matrix to wellbore as per the dual-porosity assumption is assumed.
In (C.2.10),
ft
h is the total thickness of the wellbore and fracture intersections.
For the outer boundary condition, an infinite-acting system for can be used. Then, the outer
boundary condition is given by
107

( ) ,
f fi i
p R t p p !" = = . (C.2.11)
Let us define a fracture pseudopressure by
( )
0
, 2
f
f
p
d p
f f
g
p
m p R t e dp
z
! " #
!
! $ % =
& ' (
. (C.2.12)
Then, we have
0
2 2
f
f f f
p
d p d p f f f f
f g g
m p p p
p
e dp e
R p z R z R
! " # " #
$ %
& & &
! &
' (
! = =
' (
& & & &
' (
) *
+
, (C.2.13)
and
0
2 2
f
f f f
p
d p d p f f f f
f g g
m p p p
p
e dp e
t p z t z t
! " # " #
$ %
& & &
! &
' (
! = =
' (
& & & &
' (
) *
+
. (C.2.14)
Thus, (C.2.8) through (C.2.11) become, respectively,
1 1
2
f f f
m
c fi f
m p m
R q
R R r k z t ! "
# $ %& %&
%
' = ( )
( )
* + % % %
! , (C.2.12)
( ) , 0 0
f
m R t ! = = , (C.2.13)
w
fi ft sc f
sc c
fsc
R R
k h T m
q R
p T R
!"
=
#$ % &
= '
( )
( )
#
* +
, (C.2.14)
and
( ) , 0
f
m R t ! "# = . (C.2.15)
In (C.2.12) through (C.2.15),
f f
d p
c fi c f
f
f tf g f tf g
k e k
c c
! !
"
# #
$ %
= = . (C.2.16)
and
( ) ( ) ( )
f f fi i f f
m p m p m p ! = " . (C.2.17)
108

In terms of the dimensionless variables, (4.1.2) and (4.1.4) through (4.1.6), (C.2.12) through
(C.2.15) can be written as follows:
2
2
1 fD fi ft sc f fD
D m fD
D D D sc sc fi D
m k h T L p m
R q
R R R q p T k z t
!
"
# $
% %
%
& '
( =
& '
% % %
) *
! , (C.2.18)
( ) , 0 0
fD D D
m R t = = , (C.2.19)
1
D wD
fD
D
D
R R
m
R
R
=
! " #
$ =
% &
% &
!
' (
, (C.2.20)
and
( ) , 0
fD D D
m R t !" = . (C.2.21)
In (C.2.18), following is defined
( ) ( )
f f
fi m tm g fi m tm g fi
fD
d p
f f tf g f f tf g fi
i i
c k c k
c k c k e
! " "
!
! " "
# $
= = = , (C.2.22)
It is also noted from (C.2.4) that
( )
2
2
, ,
2
4
m m
fi ft sc f fi ft sc f
am m m
m
sc sc c fi f fi sc sc m g
r r R t
k h T L p k h T p
k p p L
q
q p T k z h k q p T p z r
! !
"
=
# $
%
& '
= (
& '
& '
%
) *
! . (C.2.23)
Using (C.1.18) in (C.2.23),
( )
2
2
, ,
2
2
2 1
m m
fi ft sc f fi ft sc m
am m m
m
sc sc c fi f fi sc sc m g
r r R t
k h T L p k h T k
b p p L
q
q p T k z h k q p T p z r
! !
"
=
# $
% &
'
( ) * +
= , +
( ) * +
'
( ) - .
/ 0
! . (C.2.24)
In terms of the pseudopressure defined in (C.1.19), (C.2.24) becomes
( )
2
2
, ,
2
2
m m
fi ft sc f fi ft sc f
m m
m
r r R t
sc sc c fi f fi sc sc m
k h T L p k h T p
L k m
q
q p T k z h k q p T p r
! !
"
=
#$
=
#
! . (C.2.25)
Using (4.1.2) and (4.1.5), it is possible to write (C.2.25) in terms of dimensionless variables as
follows:
109

( ) ( )
2
, , , ,
2
2
2
D mD mD D D mD mD D
fi ft sc f f
m mD mD
m
sc sc c fi f fi m D D
r r R t r r R t
k h T L p p
Lk m m
q
q p T k z h k p r r
!
"
#
= =
$ % $ %
& &
' ( ' (
= =
' ( ' (
& &
) * ) *
!
! , (C.2.26)
where we have used the continuity of pressure at
m
r r = ; that is,
( ) ( ) , , , ,
f m m m m m
p r r R t p r r R t = = = , (C.2.27)
and defined
m
fi f
k L
k h
! =
!
. (C.2.28)
Substituting (C.2.26) into (C.2.18) and assuming
1
fi
fD fDi
fi
!
! !
!
" = = , (C.2.29)
yields the following linearized form of (C.2.29):
( ) , ,
1
2
D mD mD D
fD fD
mD
D
D D D D D
r r R t
m m
m
R
R R R r t
!
=
" # " #
$ $
$ $
% & % &
' =
% & % &
$ $ $ $
( ) ( )
!
. (C.2.30)
given as (4.3.22) in text. The Laplace transforms of (C.2.30), (C.2.20), and (C.2.21) are given,
respectively, by
( ) , ,
1
2 0
D mD mD
fD
mD
D fD
D D D D
r r R s
m
m
R sm
R R R r
!
=
" # " #
$
$ $
% & % &
' ' =
% & % &
$ $ $
( ) ( )
!
, (C.2.31)
(4.3.25) in text and where (C.2.19) is used,
1
D wD
fD
D
D
R R
m
R
R s
=
! " #
= $
% &
% &
!
' (
, (C.2.32)
and
( ) , 0
fD D
m R s !" = . (C.2.33)
(C.2.31) is given as (4.3.24) in the text.
110

Appendix C.3 Coupling Fracture and Matrix Flows to derive a Dual-Mechanism Dual-
Porosity Transfer Function
From (C.1.56), the following is given
( )
( )
( )
( ) ( )
sinh
, , , ,
sinh
mD mDi D
mD D mD fD mD bD mD
D mDi mD
r s r
m r R s m R s m R s
r s r
!
!
" # = $
% &
! ! . (C.3.1)
( )
( )
( ) ( )
, ,
1
1 coth , ,
D mD mD
mD
mDi mD mDi mD fD mD bD mD
D mD
r r R s
m
s r s r m R s m R s
r r
! !
=
" #
$
% & ' (
' ( = ) ) )
* +
* +
% &
$
, -
! ! .
(C.3.2)
If we also assume that
( ) ( ) , ,
fD mD D fD D D
m R t m R t ! ! , (C.3.3)
( ) ( ) , ,
bD mD D bD D D
m R t m R t ! ! , (C.3.4)
and substitute (C.3.2) into (C.2.31), following is obtained
( )
( ) ( )
1
2 1 coth
, , 0
fD
D mDi mD mDi mD
D D D mD
fD mD bD mD fD
m
R s r s r
R R R r
m R s m R s sm
!
" "
# $
%
%
& ' ( )
+ *
+ ,
& '
% %
- .
( ) / * * =
+ ,
!
.
(C.3.5)
Defining
( )
( )
( )
2
, 1 1 coth 1 ,
mD mDi mD mDi mD p mD
mD
f R s s r s r f R s
r s
!
" "
# $
# $ = % % %
& '
& '
!
, (C.3.6)
and
( )
( )
( )
,
,
,
bD mD
p mD
fD mD
m R s
f R s
m R s
= , (C.3.7)
(C.3.5) can be also written as
( )
1
0
fD
D fD
D D D
m
R sf s m
R R R
! "
#
#
$ %
& =
$ %
# #
' (
. (C.3.8)
111

given as (4.3.26) in text. Following the formulation of the standard dual-porosity models
(Warren and Root, 1963, de Swaan, 1976, Kazemi, 1969), lets define
( )
( )
( )
3
2 2 2
2
4 3 2
2 4 2 3
ami m ami m ami m
fi f fi m f fi f
k r k V k r
L L L
k V k r h k h
!
" # # #
!
$ % $ % & '
( ) ( ) * +
= = =
( ) ( ) * +
* +
( ) ( )
, - . / . /
. (C.3.9)
In (C.3.9), it was assumed that the flux from each matrix block is instantaneously and uniformly
distributed in one-half the fracture volume that envelopes the matrix block,
f
V
!
, (de Swaan, 1976)
and defined the shape factor, !
( )
( )
2
2
2 15
m m m
n
n V A r
!
+
= = . (C.3.10)
In the second equality of (C.3.10), we used the number of orthogonal fractures sets 3 n = for
spherical matrix blocks. Then from (C.2.28) and (C.3.9),
1
10
ami
i
b
p
!
!
" #
$ %
= +
$ %
& '
!
. (C.3.11)
Let us also define
( )
( ) ( )
( ) ( )
( ) ( )
( )
( )
3
2
4 3 2
2 4 2 3
t m t m t m
m m m
t f t m f t f
f f f
c V c r c r
c V c r h c h
! ! " !
#
! ! " !
= = = , (C.3.12)
Then from (C.1.13), (C.2.10), and (C.3.12),
( )
( )
2 2
2 2
15
fi m m tm fi m
i
mDi
mi f tf ami
i
r c k r
L c k L
! "
#
!
! " $
= = = (C.3.13)
Thus, (C.3.6) can be written as follows:
( ) ( )
15 15
, 1 1 1 coth 1 ,
5
ami
mD p mD
i
b s s
f R s f R s
s p
! " "
! !
# $
% & # $
' (
% & ) ' (* = + + + +
, -
' ( ' (
) *
. / , -
. /
. (C.3.14)
given as (4.3.27) in the text.

112

APPENDIX D
DERIVATION OF DUAL-MECHANISM DUAL-POROSITY MODEL WITH
MICROFRACTURES FOR SHALE-GAS RESERVOIRS
A composite spherical matrix block as shown in Fig. 4.2 is considered. As in the
microfracture only model the core of the matrix block is homogeneous and the surface-layer of
the matrix block has microfractures. Consistent with the conventional dual-porosity formulations
(Warren and Root, 1963, de Swaan, 1976, Kazemi, 1969), it is assumed that the pressure and flux
are uniformly distributed on the surface of each matrix block. To derive the flow solution for the
composite matrix block, the flow equations for the core and the surface layer of the matrix block
are derived and the two solutions at the boundary of the two regions are coupled.
Appendix D.1 Flow in the Matrix Core:
Mass balance for the flow of a real gas of density
g
! , viscosity
g
, and compressibility
g
c on a spherical control volume of porosity
m
! in the matrix core yields
( ) ( )
2
2
1
; 0
g rm g m mc
r v r r
r r t
! ! "
# #
$ = % %
# #
. (D.1.1)
The radial component of the effective (Darcy+slip flow) velocity in (D.1.1) is given by
am mc
prm c
g
k p
v
r
!

"
= #
"
, (D.1.2a)
where
1
am
am m
b
k k
p
! "
= +
# $
% &
. (D.1.2b)
Substituting (D.1.2a) into (D.1.1) and using the real gas equation of state,
mc g
g
g
p M
zR T
! = , (D.1.3)
following is obtained
2
2
1
; 0
am mc mc mc
c m mc
g
k p p p
r r r
r r z r t z
! "

# $ % &
% &
' ' '
( ) *+
= , , ) *
( ) *+ ) *
) *
- . ' ' '
( +
- . / 0
. (D.1.4)
113

Because
1 1 p p dz
p
z z p z dp
! "
! "
# $
% = & % # $
# $
# $
' (
' (
, (D.1.5)
1 1
g
dz
c
p z dp
= ! , (D.1.6)
1
f
c
p
!
!
"
=
"
, (D.1.7)
and
t g f
c c c = + , (D.1.8)
one can write (D.1.4) as follows:
2
2
1
1 ; 0
m tm g
am mc mc mc mc
c mc
mc g m g
c
b p p p p
r r r
r r p z r k z t
!
"

# $
% &
' ' '
( ) * +
+ = , ,
( ) * +
' ' '
( ) - .
/ 0
. (D.1.9)
Let us define a pseudopressure by
( )
0
2 1
mc
p
am
mc mc
g
b p
m p dp
p z
! "
#
$ %
# = +
$ %
#
& '
(
. (D.1.10)
Then
2
2
1 1
; 0
mc mc
mc
m
m m
r r r
r r r t !
" #
$% $% $
= & & ' (
' (
) * $ $ $
, (D.1.11)
where
( ) ( ) ( )
mc mc i i mc mc
m p m p m p ! = " (D.1.12)
and
c am
m
m tm g
k
c
!
"
#
= . (D.1.13)
A solution of (D.1.11) subject to the following conditions is required:
114

The Initial condition is given by
( ) , , 0 0
mci m i
m r R t m ! = = ! = . (D.1.14)
The boundary condition at 0 r = is
( ) ( )
0
0, ,
mc m mc
m r R t m t finite ! = = ! = . (D.1.15)
The boundary condition at the interface between the core and the surface layer of the matrix block,
mc
r , is obtained from the continuity of the pressures given by
( ) ( ) ( ) , , , , , ,
mc mc mc m f mf mc m mb mf mc m
m p r r R t m p r r R t m p r r R t ! ! ! " # " # " # = = = $ =
% & % & % &
, (D.1.16)
where we have defined
( )
( ) , ,
, , 2 1 1
i
mf mf
mf mc m
p
d p d p
am
mb mf mc m
p r R t
g
b p
m p r r R t e e dp
p z
! ! " # "
$ % & '
!
( ) *+
! $ % " = = # +
, -
( ) *+
!
. / , -
0
. (D.1.17)
For convenience, the following dimensionless variables are defined:
, where , , , or
f ft sc
D
sc sc
k h T
m m f mc mf ms
q p T
! !
"
! = # = , (D.1.18)
2
fi
D
t t
L
!
=
, (D.1.19a)
where L is a characteristic length,
( )
fi
fi
f tf g
i
k
c
!
"
= , (D.1.19b)
and k
fi
, is the initial value of the stress dependent macrofracture permeability defined by
( )
f i f
d p p
f fi
k k e
! !
= , (D.1.19c)
D
r
r
L
= , (D.1.20)
and
115

D
R
R
L
= . (D.1.21)
Using the dimensionless variables, (D.1.11) and (D.1.14) through (D.1.16) can be written as
2
2
1 1
; 0
mcD mcD
D D mcD
D D D mD D
m m
r r r
r r r t !
" #
$ $ $
% &
= ' '
% &
$ $ $
( )
, (D.1.22)
where
( )
f tf g am
m i
mD
fi m tm g fi
c k
c k
!
"
"
" !
= = , (D.1.23)
( ) , , 0 0
mcD D mD D
m r R t = = , (D.1.24)
( ) ( )
0
0, ,
mcD D mD D mcD D
m r R t m t finite = = = , (D.1.25)
and
( ) ( ) ( ) , , , , , ,
mcD D mcD mD D mfD D mcD mD D mbD D mcD mD D
m r r R t m r r R t m r r R t = = = ! = . (D.1.26)
To proceed with an analytical solution, it is assumed that
( )
( )
f tf g m
mi i
mD mDi
fi m tm g fi
i
c k
c k
!
"
" "
" !
# = = (D.1.27)
and linearize (D.1.22) as follows:
2
2
1 1
; 0
mcD mcD
D D mcD
D D D mDi D
m m
r r r
r r r t !
" #
$ $ $
% &
= ' '
% &
$ $ $
( )
. (D.1.28)
Lets consider the substitution
( ) ( ) , , , ,
mcD D mD D D mcD D mD D
w r R t r m r R t = . (D.1.29)
Then, (D.1.28) and (D.1.24) through (D.1.26) can be expressed in terms of
( ) , ,
mcD D mD D
w r R t ,
respectively, as follows:
2
2
1
mcD mcD
D mDi D
w w
r t !
" "
=
" "
, (D.1.30)
116

( ) , , 0 0
mcD D mD D
w r R t = = , (D.1.31)
( ) 0, , 0
mcD D mD D
w r R t = = , (D.1.32)
and
( ) ( ) ( ) , , , , , ,
mcD mcD mD D mfD mcD mD D mbD mcD mD D
w r R t w r R t w r R t = ! . (D.1.33)
If the Laplace Transformation of (D.1.30), (D.1.32), and (D.1.33) is taken, following is obtained,
respectively,
2
2
0
mcD
mcD
D mDi
w s
w
r !
"
# =
"
, (D.1.34)
where (D.1.31) is used,
( ) 0, , 0
mcD D mD
w r R s = = (D.1.35)
and
( ) ( ) ( ) , , , , , ,
mcD mcD mD mfD mcD mD mbD mcD mD
w r R s w r R s w r R s = ! . (D.1.36)
The solution of (D.1.34) is given by
( ) ( )
exp exp
mcD mDi D mDi D
w A s r B s r ! ! = " + . (D.1.37)
From (D.1.35), A B = ! . Then,
( ) ( ) ( )
exp exp sinh
mcD mDi D mDi D mDi D
w B s r s r C s r ! ! !
" #
= $ $ =
% &
. (D.1.38)
Using (D.1.35),
( ) ( )
( )
, , , ,
sinh
mfD mcD mD mbD mcD mD
mDi mcD
w r R s w r R s
C
s r !
"
= . (D.1.39)
Thus,
( )
( )
( )
( ) ( )
sinh
, , , , , ,
sinh
mDi D
mcD D mD mfD mcD mD bD mcD mD
mDi mcD
s r
w r R s w r R s w r R s
s r
!
!
" # = $
% &
. (D.1.40)
Finally, using (D.1.29),
117

( )
( )
( )
( ) ( )
sinh
, , , , , ,
sinh
mcD mDi D
mcD D mD mfD mcD mD mbD mcD mD
D mDi mcD
r s r
m r R s m r R s m r R s
r s r
!
!
" # = $
% &
. (D.1.41)
Appendix D.2 Flow in the Matrix Surface Layer:
To model flow in the matrix surface layer, the transient dual-porosity idealization for a
naturally fractured porous medium suggested by de Swaan (1976) and Kazemi (1969) is used.
Assuming that the thickness of the surface layer (
ms m mc
h r r = ! ) is small compared to the radius of
the matrix core (
mc
r ) and the pressure and flux are uniformly distributed on the boundaries (
mc
r
and
m
r ) of the surface layer, the surface layer can be approximated by a linear system as sketched
in Figs. 4.3 and 4.4.
Flow in the Surface-Layer Microfractures: For the geometry considered in Figs. 4.3 and
4.4, one-dimensional flow in the matrix fractures is described by
( ) ( ) ( )
, ;
g rmf g ms g mf mc m
v q r t r r r
r t
! ! ! "
# #
$ + = % %
# #
! , (D.2.1)
where
( ) ,
ms
q r t ! is the influx from matrix per unit volume of fracture in the matrix surface layer
per unit time. Following de Swaan O (1976), it was assumed that the flux from each surface of
the matrix slab is instantaneously and uniformly distributed in one-half the fracture slab adjacent
to the matrix surface. If it is assumed that
f mf mf
V A h =
!
denote the volume of each fracture slab,
where
mf
A is the area of the contact surface between the matrix and fracture, and
( ) 2,
ms mm
q h t ! = denote the matrix efflux per unit time,
( ) ( )
( ) ( ) ( )
( ) 2,
, 2, 2 ,
, ,
2
mm
g ms mm g
am ms
g ms c
mf mf mf g
h t
r t q h t r t
k p
r t q r t
A h h
!
" ! "
" #
!
=
$ % & '
=
(
) * + ,
= = -
) * + ,
) *
(
+ ,
. / 0 1
! . (D.2.2)
Using (D.2.2) and
mf mf
rmf c
g
k p
v
r
!

"
= #
"
, (D.2.3)
where
mf mf
d p
mf mfi
k k e
! "
= , (D.2.4)
118

and
mf g
g
g
p M
zR T
! = , (D.2.5)
one can write (D.2.1) as
( ) 2,
2
;
mf mf
mm
d p mf mf am mf mf tmf g mf mf
ms
mc m
h t g mfi mf g c mfi g
p p k p c p p
p
e r r r
r z r k h z k z t
!
"
! #
$ %
=
& '
& ' ( (
( (
) *
$ = + + ) *
) * ) *
) *
, - ( ( ( (
, -
,(D.2.6)
where (D.1.5) through (D.1.8) have been used. The initial condition is given by
( ) , 0
mf i
p r t p = = . (D.2.7)
One of the boundary conditions is the continuity of flux at the interface of the matrix core and the
matrix surface layer,
mc
r ; that is,
( )
( )
( )
( ) , ,
mc mc
g mf g mc
r t r t
q q ! ! = ! ! . Using the bulk fracture
permeability,
( )
mf mf mf mm mf mf mf mm
k k h h h k h h = + !
!
, (D.1.3), and (D.2.4), this condition is given
by
( ) ( ) , ,
mc mc
mf mf mf
mc mc
mf am
mm g g
r t r t
h p p
p p
k k
h z r z r
! " ! "
#
#
$ % $ %
=
$ % $ %
$ % $ %
# #
& ' & '
. (D.2.8)
The other condition is the continuity of pressure at the matrix-fracture interface at,
m
r , given by
( ) ( ) , ,
mf m f m
p r r t p R R t = = = . (D.2.9)
In terms of the pseudopressure defined by
( )
0
2
mf
mf
p
d p
mf mf
g
p
m p e dp
z
! " #
!
! =
$
(D.2.10)
(D.2.6) through (D.2.49) are written as follows, respectively:
( ) 2,
2 1
;
mm
mf mf
am ms
mc m
h t mfi mf mf
m m
k m
r r r
r r k h t
!
! "
=
# $ # $ %& %&
%& %
' = ( ( ) * ) *
) * ) *
+ , + , % % % %
, (D.2.11)
( ) , 0 0
mf
m r t ! = = , (D.2.12)
119

( ) ( ) , ,
mc mc
mf mf
mc
mfi am
r t r t mm
h m
m
k k
h r r
! " ! " #$
#$
= % & % &
% & % &
' ( ' ( # #
, (D.2.13)
and
( ) ( ) ( ) , , ,
mf m f m b m
m r r t m R R t m R R t ! = = ! = "! = . (D.2.14)
In (D.2.11) through (D.2.14),
( ) ( ) ( )
mf mf i i mf mf
m p m p m p ! = " , (D.2.15)
( ) ( ) ( )
f f i i f f
m p m p m p ! = " , (D.2.16)
( )
( )
( ) ,
, 2
i
f mf
f m
p
d p d p
b m
p R t
g
p
m R R t e e dp
z
! ! " # #
!
! # = = "
$
. (D.2.17)
and
c mf
mf
mf tmf g
k
c
!
"
#
= . (D.2.18)
In terms of the dimensionless variables,
( )
2
2
1,
1
;
3
D D
mfD mfD
m msD
mcD D mD
D D mfD D
t
m m
m
r r r
r t
!
"
! #
=
$ %
& &
&
' (
) = * *
' (
& & &
+ ,
, (D.2.19)
where
( )
f tf g mf
mf
i
mfD
fi mf tmf g fi
c k
c k
!
"
"
" !
= = , (D.2.20a)
and, Following Serra et al. (1983), following is defined
2
2
12
am mm
m
mm mfi mf
k h L
h k h
!
" #
$ %
=
$ %
$ %
& '
(D.2.20b)
To linearize (D.2.19), it is assumed that
( )
mfi
mf mfi
mf tmf g
i
k
c
! !
"
# = , (D.2.21)
120

and define
mfi
mfD mfDi
fi
!
! !
!
" = . (D.2.22)
Then, (D.2.19), and (D.2.11) through (D.2.14) become, respectively,
( )
2
2
1,
1
;
3
D D
mfD mfD
m msD
mcD D mD
D D mfDi D
t
m m
m
r r r
r t
!
"
! #
=
$ %
& &
&
' (
) = * *
' (
& & &
+ ,
. (D.2.23)
( ) , 0 0
mfD D D
m r t = = , (D.2.24)
( ) ( )
2
2
, ,
12
mcD D mcD D
mfD
m mm mcD
D D
r t r t
m
h m
r L r
!
" # " #
$
$
% & % &
=
% & % &
$ $
' ( ' (
, (D.2.25)
and
( ) ( ) ( ) , , ,
mfD mD D fD mD D bD mD D
m r t m R t m R t = ! . (D.2.26)
Flow in the Surface-Layer Matrix: Based on the representation of the surface-layer as
slabs of matrix and fracture media (Fig. 4.4), one-dimensional flow in the surface-layer matrix is
defined by
( ) ( )
; 0 2
g m g m mm
v h
t
!
" " # !
!
$ $
% = & &
$ $
. (D.2.27)
Substituting
am ms
m c
g
k p
v
!
"
!
#
= $
#
, (D.2.28)
and
ms g
g
g
p M
zR T
! = , (D.2.29)
one can write (D.2.27) as
1 ; 0 2
m tm g
am ms ms ms ms
mm
ms g c m g
c
b p p p p
h
p z k z t
!
"
" " #
$ %
& '
( ( (
) * + ,
+ = - -
) * + ,
( ( (
) * . /
0 1
. (D.2.30)
121

The initial and boundary conditions are given by
( ) , 0
ms i
p t p ! = = , (D.2.31)
( ) 0,
0
ms
t
p
!
!
=
"
=
"
, (D.2.32)
( ) ( ) 2, ,
ms mm mf
p h t p r t ! = = , (D.2.33)
In terms of the pseudopressure defined by (D.1.10), (D.2.29) through (D.2.33) are written as
follows, respectively:
2
2
1
; 0 2
ms ms
mm
m
m m
h
t
!
! "
# $ #$
= % %
# #
, (D.2.34)
( ) , 0 0
ms
m t ! " = = , (D.2.35)
( ) 0,
0
ms
t
m
!
!
=
"#
=
"
, (D.2.36)
and
( ) ( ) ( ) 2, , ,
ms mm mf bs
m h t m r t m r t ! " = = " #" , (D.2.37)
where
( )
( ) ,
, 2 1 1
i
mf mf
mf
p
d p d p
am
bs
p r t
g
b p
m r t e e dp
p z
! ! " # "
$ % & '
!
( ) *+
! " = # +
( ) *+
, - . /
0
. (D.2.38)
( ) ( ) ( )
ms ms i i ms ms
m p m p m p ! = " . (D.2.39)
Defining
2
D
mm
h
!
! = , (D.2.40)
and using the dimensionless variables, following is obtained
2 2
2 2
; 0 1
4
msD mm msD
D
D mD D
m h m
L t
!
! "
# #
= $ $
# #
, (D.2.41)
122

( ) , 0 0
msD D D
m t ! = = , (D.2.42)
( ) 0,
0
D D
msD
D
t
m
!
!
=
"
=
"
, (D.2.43)
and
( ) ( ) ( ) 1, , ,
msD D D mfD D D bsD D D
m t m r t m r t ! = = " . (D.2.44)
To proceed with an analytical solution, it is assumed that
mD mDi
! ! " (D.2.45)
and define
( )
( )
f tf g mfi
mfi
i
mfDi
fi mf tmf g fi
i
c k
c k
!
"
"
" !
= = . (D.2.46)
Following Serra et al. (1983), it is also defined that
( )
( )
m tm mm
i
m
mf tmf mf
i
c h
c h
!
"
!
= . (D.2.47)
Then, (82) becomes
2
2
3
; 0 1
msD m msD
D
D fmDi m D
m m
t
!
"
" # $
% %
= & &
% %
. (D.2.48)
Taking the Laplace transformation of (D.2.48), (D.2.43), and (D.2.44), respectively,
2
2
3
=0
msD m
msD
D m mfDi
m
sm
!
" # $
%
&
%
, (D.2.49)
where (D.2.42) is used,
( ) 0,
0
D
msD
D
s
m
!
!
=
"
=
"
, (D.2.50)
and
123

( ) ( ) ( ) 1, , ,
msD D mfD D bsD D
m s m r s m r s ! = = " . (D.2.51)
The general solution of (D.2.49) is given by
3 3
exp exp
m m
msD D D
m mfDi m mfDi
m A s B s
! !
" "
# $ # $
% & % &
' ( ' (
= ) +
' ( ' (
' ( ' (
* + * +
. (D.2.52)
Using (D.2.50) and (D.2.51), the solution is obtained as
( ) ( )
3
cosh
, ,
3
cosh
m
D
m mfDi
msD mfD D bsD D
m
m mfDi
s
m m r s m r s
s
!
"
# $
!
# $
% &
' (
' (
' (
) *
+ , = -
. /
% &
' (
' (
' (
) *
. (D.2.53)
Coupled Flow Solution for the Matrix Surface Layer: Taking the Laplace transformation
of (D.2.23), following is obtained
( )
2
2
1,
0
3
D
mfD
m msD
mfD
D D mfDi
s
m
m s
m
r
!
"
! #
=
$ %
&
&
' (
) ) =
' (
& &
* +
, (D.2.54)
where (D.2.24) is used. From (D.2.53)
( )
( ) ( )
1,
3 3
tanh , ,
D
sD m m
mfD D bsD D
D m mfDi m mfDi
s
m
s s m r s m r s
!
" "
! # $ # $
=
% &
% &
'
( )
( )
* + = , ,
- .
( )
( )
( )
'
/ 0
/ 0
. (D.2.55)
Substituting into (D.2.54), following is obtained
2
2
0
mfD
m mfD
D
m
u m
r
!
" =
!
. (D.2.56)
In (D.2.52), following is defined
( )
m m
u sf s = , (D.2.57)
where
( )
( )
, 3 1
1 tanh 1
3 ,
m m mfDi
bsD D m
m
mfDi m mfDi mfD D
m r s
f s
s m r s
! " #
"
# ! #
$ % & ' ( )
* *
+ ,
- .
= / /
0 1
+ ,
- .
+ , * *
- .
2 3 4 5 6 7
, (D.2.58)
124

given as (6.2.4) in text. Taking the Laplace transform of (D.2.25),
( )
( )
,
,
mcD
mfD
mf mfD mcD
D
r s
m
f m r s
r
! "
#
$ %
=
$ %
#
& '
, (D.2.59)
where, from (D.1.41), following is used
( )
( )
( )
( )
( )
,
, 1
coth 1 ,
,
mcD
mbD mcD mcD
mDi mDi mcD mfD mcD
D mcD mfD mcD
r s
m r s m
s s r m r s
r r m r s
! !
" #
" # $ %
&
' (
) * ' (
= + +
' (
) * ' (
&
' ( , - . /
. /
, (D.2.60)
and defined
( )
2
2
coth 1
12
mm m
mf mDi mcD mDi mcD
mcD
h
f s r s r
r L
!
" "
# $
= %
& '
. (D.2.61)
given as (6.2.6) in text. Taking the Laplace transform of (D.2.26),
( ) ( ) ( ) , , ,
mfD mD fD mD bD mD
m r s m R s m R s = ! . (D.2.62)
The general solution of (D.2.56) is given by
( ) ( )
exp exp
mfD m D m D
m A u r B u r = ! + . (D.2.63)
From (D.2.59) and (D.2.63),
( ) ( ) ( ) ( )
exp exp exp exp
m m mcD m m mcD mf m mcD m mcD
A u u r B u u r f A u r B u r
! "
# # + = # +
$ %
,
(D.2.64)
which yields
( )
( )
( )
exp 2
m mf
m mcD
m mf
u f
B A u r
u f
+
= !
!
. (D.2.65)
Substituting (D.2.65) into (D.2.63), following is obtained
( )
( )
( )
( )
1 exp 2 exp
m mf
mfD m D mcD m D
m mf
u f
m u r r A u r
u f
! "
+
# #
$ %
= + & &
' (
) *
# #
&
+ ,
. (D.2.66)
Using (D.2.62) and (D.2.66),
125

( )
( )
( )
( )
( )
( )
( )
exp
,
1 ,
,
1 exp 2
m mD
bD mD
fD mD
fD mD m mf
m mD mcD
m mf
u r
m R s
A m R s
m R s u f
u r r
u f
! "
# $
= %
# $
& '
+
# $
( ) * *
! "
+ %
+ ,
( )
* * %
- .
. (D.2.67)
Substituting (D.2.67) into (D.2.66), the particular solution is obtained as follows:
( )
( ) ( )
( )
( ) ( )
( )
( )
( )
( )
( )
exp 2
, exp
exp 2
,
1 ,
,
m mf m mf m D mcD
mfD D m mD D
m mf m mf m mD mcD
bD mD
fD mD
fD mD
u f u f u r r
m r s u r r
u f u f u r r
m R s
m R s
m R s
! "
# + + #
$ %
! "
= #
$ %
! "
# + + #
$ %
! "
& '
( #
& '
& '
$ %
. (D.2.68)
Appendix D.3 Radial Dual-Porosity Model with Transient Flow in Spherical Matrix Blocks:
Consider a cylindrical, dual-porosity reservoir. Mass balance for flow of a real gas of
density
g
! , viscosity
g
, and compressibility
g
c on a fracture control volume of porosity
f
!
yields
( ) ( ) ( )
1
,
g Rf g mf g f
R v q R t
R R t
! ! ! "
# #
$ + =
# #
! , (D.3.1)
where
( ) ,
mf
q R t ! is the influx from matrix per unit volume of fracture per unit time. Note that fluid
transfer from matrix to fractures is assumed to be through the matrix fractures only; that is, when
flux from the matrix is considered, only the flux from matrix fractures is considered. Following
de Swaan O (1976), it was assumed that the flux from each matrix block is instantaneously and
uniformly distributed in one-half the fracture volume that envelopes the matrix block,
f
V
!
. If we
let
( ) , ,
mf m m
q r r R t = denote the matrix outflux per unit time,
( ) ( )
( ) ( )
( )
( )
( )
2
, , , ,
, ,
4
, ,
2 2
m m m m
g mf g mf
g mf c mf
f f g
r r R t r r R t
R t q R t p
R t q R t r k
V V r
! !
"
! #

= =
$ % & '
(
) * + ,
= = -
) * + ,
) *
(
+ ,
. / 0 1
!
!
! !
,(D.3.2)
where
mf
k
!
is the bulk permeability of the fracture matrix defined by (see Fig. 4.4)
mf
mf mf
mm
h
k k
h
=
!
. (D.3.3)
126

If it is assumed that the fracture volume that envelopes the matrix block,
f
V
!
, has a uniform
average thickness of
f
h , and that
f
h is small compared with the dimensions of the matrix block,
m
r , it can approximated that
2
1
4
2 2 2
f f
f m m
h h
V A r ! " =
!
. (D.3.4)
Using (D.3.3) and (D.3.4), (D.3.2) becomes
( ) ( )
( ) , ,
2
, ,
m m
mf g mf
g mf c mf
f mm g
r r R t
h p
R t q R t k
h h r
!
! "

=
# $
%
& '
= (
& '
& '
%
) *
! . (D.3.5)
The radial component of the velocity vector for flow in fractures is given by
f f
Rf c
g
k p
v
R
!

"
= #
"
. (D.3.6)
Substituting (D.3.5), (D.3.6) and the real gas equation of state,
f g
g
g
p M
zR T
! = , (D.3.7)
into (D.3.1), following is obtained
( ) , ,
2
1
f f
m m
d p f f mf mf f mf f tf f f
g fi f mm g c fi
r r R t
p p k h p p c p p
e R
R R z R k h h z r k z t
!
"
# $
=
% & % &
' ' '
'
( ) ( )
# =
( ) ( )
( ) ( )
' ' ' '
* + * +
. (D.3.8)
Assume that initially
( ) , 0
f i
p R t p = = . (D.3.9)
For a vertical, fully penetrating well of radius
w
R located at the center of the reservoir, 0 R = , the
inner boundary condition is
2 2
w w
g g f f f ft sc f f
sc ft c c
gsc gsc g fsc g
R R R R
k p k h T p p
q q h R R
R p T z R
! !
" # "#
! !
= =
$ % $ % & &
= = = ' ( ' (
' ( ' (
& &
) * ) *
, (D.3.10)
127

where production at a constant surface rate of,
sc
q , is considered, ignored the effects of wellbore
storage, and no flow from matrix to wellbore as per the dual-porosity assumption is assumed. In
(D.3.10),
ft
h is the total thickness of the wellbore and fracture intersections.
For the outer boundary condition, one can use an infinite-acting system. Then, the outer boundary
condition is given by
( ) ,
f i
p R t p !" = . (D.3.11)
In terms of the pseudopressure (D.3.8)-(D.3.11) are given, respectively, by
( ) , ,
2
1 1
m m
f mfi mf mf f
r r R t fi f mm f
m k h m m
R
R R R k h h r t !
=
" # " # $% $% $%
$
& = ' ( ' (
' ( ' (
) * ) * $ $ $ $
, (D.3.12)
( ) , 0 0
f
m R t ! = = , (D.3.13)
w
fi ft sc f
sc
fsc
R R
k h T m
q R
p T R
!
=
"# $ %
= &
' (
' (
"
) *
, (D.3.14)
and
( ) , 0
f
m R t ! "# = . (D.3.15)
In (D.3.12), following is defined
c f
f
f tf g
k
c
!
"
#
= . (D.3.16)
In terms of dimensionless variables, one can write (D.3.12) through (D.3.15) as follows,
respectively:
( ) , ,
2
1
D mD mD D
fD mfi mf mfD fD
D
D D D fi f mm D D
r r R t
m k h L m m
R
R R R k h h r t
=
! " ! "
# # #
#
$ % $ %
& =
$ % $ %
# # # #
' ( ' (
, (D.3.17)
( ) , 0 0
fD D D
m R t = = , (D.3.18)
1
D wD
fD
D
D
R R
m
R
R
=
! " #
$ =
% &
% &
!
' (
, (D.3.19)
and
128

( ) , 0
fD D D
m R t !" = . (D.3.20)
To linearize (D.3.17), it was also assumed
( )
1
f tf g f
f
i
fD
fi f tf g fi
c k
c k
!
"
"
" !
= = # . (D.3.21)
Taking the Laplace transform of (D.3.17), (D.3.19) and (D.3.20), following is obtained,
respectively,
( ) , ,
2
1
0
mD mD
fD mfi mf mfD
D fD
D D D fi f mm D
r R s
m k h L m
R sm
R R R k h h r
! " ! "
# #
#
$ % $ %
& & =
$ % $ %
# # #
' ( ' (
, (D.3.22)
where (D.3.18) is used,
1
D wD
fD
D
D
R R
m
R
R s
=
! " #
= $
% &
% &
!
' (
, (D.3.23)
and
( ) , 0
fD D
m R s !" = . (D.3.24)
From (D.2.68),
( )
( )
, ,
5
,
mD mD
mfD
mm
f m fD mD
D mf mD
r R s
m
h
f f sm R s
r h r
! "
#
$ %
= &
$ %
#
' (
, (D.3.25)
where
( ) ( )
( )
( ) ( )
( )
( )
( )
exp 2
,
1
5 exp 2 ,
m mf m mf m mD mcD
mf mD bD mD
f
m mm m mf m mf m mD mcD fD mD
u f u f u r r
h r m R s
f
u h u f u f u r r m R s
! "
# $ % &
# $
' ' + '
( ( ) *
+ , - .
= '
/ 0
+ , - .
( ( # $
' + + '
+ , 1 2 ) * ) * 3 4
.(D.3.26)
given as (6.2.5) in text. One can now write (D.3.22) as follows:
2
2
10
1
1 0
fD mfi m
D f m fD
D D D fi f m
m k r L
R f f sm
R R R k h r
! "
! "
#
#
$ %
$ %
+ & =
$ %
$ %
$ %
# #
' (
' (
. (D.3.27)
Following the formulation of the standard dual-porosity models (Warren and Root, 1963, de
Swaan, 1976, Kazemi, 1969), lets define
129

( )
( )
( )
3 2
2 2 2 2
2 2 2
4 3 2 2
15
10
2 4 2 3 3
mfi m mfi m mfi m mfi m mfi m
fi f fi m f fi f m fi f fi f m
k V k r k r k r k r L
L L L L
k V k r h k h r k h k h r
!
" # # #
!
$ % $ % & ' & '
( ) ( ) * + * +
= = = = =
( ) ( ) * + * +
* + * +
( ) ( )
, - , - . / . /
.
(D.3.28)
In (D.3.28), it was assumed that the flux from each matrix block is instantaneously and uniformly
distributed in one-half the fracture volume that envelopes the matrix block,
f
V
!
, (de Swaan, 1976)
and the shape factor, ! , is defined by
( )
( )
2
2
2 15
m m m
n
n V A r
!
+
= = . (D.3.29)
In the second equality of (D.3.29), the number of orthogonal fractures sets 3 n = for spherical
matrix blocks is used. Then from (D.3.27) and (D.3.28),
1
0
fD
D fD
D D D
m
R um
R R R
! "
#
#
$ %
& =
$ %
# #
' (
, (D.3.30)
given as (6.2.1) in text where
( ) u sf s = , (D.3.31)
and
( ) 1
f m
f s f f ! = " . (D.3.32)
The general solution of (D.3.30) is given by
( ) ( ) 0 0 fD D D
m AK uR BI uR = + , (D.3.33)
and, from (D.3.23) and (D.3.24), the particular solution is obtained as
( )
( )
0
1
D
fD
wD wD
K uR
m
s uR K uR
= . (D.3.34)
Note that, in the above derivations, one can substitute the following relations:
( )
( )
2 2
15 12
t g m
fi
mi mD m mmD
mDi
fi t g fi m
mi
c k
r h
c k
!
" # #
"
" ! $ $
= = = (D.3.35)
130

and
2
15
mfi
mD
mfDi
fi
r
!
"
!
! #
= = . (D.3.36)
where
( )
( ) ( )
( ) ( )
( ) ( )
( )
( )
3
2
4 3 2
2 4 2 3
t m t m t m
mfi mfi mfi
t f t m f t f
fi fi fi
c V c r c r
c V c r h c h
! ! " !
#
! ! " !
= = = (D.3.37)

Você também pode gostar