You are on page 1of 219

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2012-10-03

Geomechanical Productivity and Injectivity Modeling


of Multifractured Horizontal Wells

Islam, Arshad

Islam, A. (2012). Geomechanical Productivity and Injectivity Modeling of Multifractured Horizontal


Wells (Unpublished master's thesis). University of Calgary, Calgary, AB.
doi:10.11575/PRISM/26287
http://hdl.handle.net/11023/275
master thesis

University of Calgary graduate students retain copyright ownership and moral rights for their
thesis. You may use this material in any way that is permitted by the Copyright Act or through
licensing that has been assigned to the document. For uses that are not allowable under
copyright legislation or licensing, you are required to seek permission.
Downloaded from PRISM: https://prism.ucalgary.ca
UNIVERSITY OF CALGARY

Geomechanical Productivity and Injectivity Modeling of Multifractured Horizontal Wells

by

Arshad Islam

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

DEPARTMENT OF CHEMICAL AND PETROLEUM ENGINEERING

CALGARY, ALBERTA

SEPTEMBER, 2012

© Arshad Islam 2012


Abstract

The advances in hydraulic fracturing technology and horizontal well completions have

led in recent years to rapid rise in exploitation and development of tight gas and shale

plays all over the world, and particularly in North America. The popularity of new field

technology has in fact raised many new questions. In particular, for forecasting the

productivity and EUR of multifractured horizontal wells, it is not clear if conventional

reservoir simulation concepts can be adapted for modeling or if extra physics must be

included to obtain realistic solutions.

This work presents various methods to model multifractured horizontal wells in tight gas

sands using a conventional reservoir simulator coupled with geomechanics. Two actual

wells in the same formation but fractured with different techniques (i.e., X-link gelled

water fracs and un-gelled water (slick water) fracs) are studied. Detailed investigation of

the role of fracture conductivity, effects of initial permeability level, net pay thickness,

assumed size of the stimulated reservoir volume (SRV), pressure or stress dependent

permeability of the SRV and formation, and virgin reservoir were carried out by history

matching the rate and cumulative production. It was established that i) history match is

not possible without use of stress or pressure dependent permeability and ii) permeability

dependence on pressure inside stimulated reservoir volume must be larger than in the rest

of the formation. It was also observed that the standard method for using the same

geomechanical data both in uncoupled reservoir and coupled geomechanical model will

give incorrect results in terms of production.

A new method based on uniaxial deformation theory is proposed to more accurately

ii
approximate the geomechanical effects in conventional reservoir simulators without

running a fully coupled simulator. The production results from the uncoupled reservoir

modeling using the new method for correcting the permeability data for poroelastic

effects were remarkably similar to rigorously coupled geomechanical modeling. This

work will be of importance to engineers in analyzing and forecasting production

performance of multifractured horizontal completions using numerical models. It will

allow engineers to use uncoupled (conventional) reservoir modeling as a practical

approximation of more complex coupled geomechanical models.

Same wells were used to model the injection process. History matching of the field

bottomhole injection pressure using uncoupled and fully coupled geomechanical models

showed the importance of including the geomechanics. Sensitivity study of several

history matching parameters such as fracture permeability reduction factor, limiting

length of fracture propagation and Biot’s constant was performed and its effects on

injection pressure were discussed. Possibility of shear failure in the SRV during injection

was also studied, using the Mohr-Coulomb failure criterion. No shear failure was

detected when intact rock shear strength parameters were used, but significant shear

regions were generated when failure envelope represented fractured or weakened rock.

The main contributions of this work are i) better understanding of the role of the

geomechanical effects in both production and injection modeling, ii) the demonstration of

the need for coupled geomechanical modeling in injection, and iii) presentation of

techniques for approximating these effects in uncoupled reservoir simulation. In addition,

valuable insights were gained into the mechanics of fracturing and reservoir behavior

during production.

iii
Acknowledgements

I wish to express my sincere appreciations from the bottom of my heart to all those who

supported me to complete this thesis. My deepest appreciation goes to my supervisor, Dr.

Antonin Settari for his continuous support, patience, encouragement and insightful

guidance throughout the whole study and research. I would like to thank him for his trust

in me and giving me the opportunity to work with him. I have enjoyed every single

moment of working with him and years of studies and research under his guidance.

I would like to thank Dr. Roberto Aguilera, Dr. Christopher Clarkson, and Mohammad

Ali Bagheri for being the members of the supervisory and examination committee of

MSc. Program. I am grateful to management of Apache Petroleum Corporation for

providing us field and research data. I also wish to thank Mr. Vikram Sen for helping us

and Taurus Reservoir Solutions Ltd. for the generous donation of the TRS® and GeoSim®

software and University of Calgary, the Department of Chemical and Petroleum

Engineering and Graduate Studies at the University of Calgary for providing excellent

research environment and facilities. Last but not the least my colleagues in my research

group and Ayaz Mehmood, Faisal Iqbal, Faraz Rasheed, Omair Shafiq, Syed Saad Alam,

and Wajih Naeem – Thank you all for your support, advice, friendship and making my

stay memorable and enjoyable.

Finally I would like to express my gratitude to my family for their endless understanding,

support, encouragement and love during the past and future tough and easy times. I am

really thankful to my Father who taught me how to live and how to be strong on the

curvy road of life and my passionate loving Mother.

iv
Dedication

To my Father, Muhammad Islam and my Mother, Rabia Islam,

for their years of love, affection, support and encouragement.

v
Table of Contents

Abstract ............................................................................................................................... ii
Acknowledgements ............................................................................................................ iv
Dedication ............................................................................................................................v
Table of Contents ............................................................................................................... vi
List of Tables .......................................................................................................................x
List of Figures and Illustrations ........................................................................................ xii
List of Symbols, Abbreviations and Nomenclature ......................................................... xvi

CHAPTER ONE: INTRODUCTION ..................................................................................1


1.1 BACKGROUND .......................................................................................................1
1.2 RESEARCH OBJECTIVES ......................................................................................3
1.2.1 Production Modeling of Multifractured Horizontal Wells ................................3
1.2.2 Injection Modeling of Multifractured Horizontal Wells ...................................4
1.3 THESIS ORGANIZATION ......................................................................................5

CHAPTER TWO: LITERATURE REVIEW ......................................................................8


2.1 PRESSURE / STRESS DEPENDENT ROCK PROPERTIES .................................9
2.1.1 Experimental Work ...........................................................................................9
2.1.2 Mathematical Correlations ..............................................................................13
2.2 EFFECTS OF PRESSURE / STRESS DEPENDENT PERMEABILITY ON
GAS PRODUCTION .............................................................................................20
2.3 CONVENTIONAL FRACTURE MODELS ...........................................................23
2.3.1 Two – Dimensional (2-D) Fracture Models ....................................................24
2.3.1.1 PKN Fracture Model ..............................................................................24
2.3.1.2 KGD Fracture Model .............................................................................26
2.3.2 Three – Dimensional (3-D) Fracture Models ..................................................27
2.4 UNCONVENTIONAL FRACTURE MODELS .....................................................28
2.5 SUMMARY .............................................................................................................31

vi
CHAPTER THREE: SUMMARY OF THE FIELD DATA .............................................33
3.1 BASIC DATA..........................................................................................................33
3.2 STIMULATION DATA ..........................................................................................36
3.3 PRODUCTION DATA............................................................................................39
3.4 GEOMECHANICAL DATA ..................................................................................41

CHAPTER FOUR: MODELING OF STATIC FRACTURES .........................................45


4.1 REPRESENTATION OF FRACTURES IN RESERVOIR SIMULATOR ............47
4.1.1 Transmissibility Modifications ........................................................................47
4.1.2 Element of Symmetry Model ..........................................................................50
4.1.3 Grid Refinement ..............................................................................................51
4.2 FRACTURE DENSITY ..........................................................................................51
4.3 FRACTURES PARAMETERS ESTIMATION .....................................................53
4.4 SUMMARY .............................................................................................................57

CHAPTER FIVE: HISTORY MATCHING USING PSEUDO CONTINUUM


APPROACH – UNCOUPLED MODELS................................................................59
5.1 ROLE OF FRACTURE CONDUCTIVITY ............................................................61
5.1.1 Baseline Propped Fracture Conductivity .........................................................62
5.1.2 Realistic Propped Fracture Conductivity.........................................................63
5.1.3 Infinite Fracture Conductivity – An Analytical Approach ..............................67
5.2 EFFECT OF NET PAY THICKNESS (NPT) .........................................................71
5.3 EFFECTS OF STIMULATED RESERVOIR VOLUME (SRV) ...........................73
5.4 HISTORY MATCHING USING MANUAL CHANAGE OF RESERVOIR
PERMEABILITY WITH TIME ............................................................................81
5.5 SUMMARY .............................................................................................................84

CHAPTER SIX: HISTORY MATCHING AND MODELING OF


GEOMECHANICAL EFFECTS USING PRESSURE / STRESS DEPENDENT
PERMEABILITY .....................................................................................................85
6.1 DISCUSSION ON NEW INFORMATION ............................................................88
6.2 PRESSURE / STRESS DEPENDENT PERMEABILITY CORRELATION ........91

vii
6.3 HISTORY MATCHING USING PRESSURE DEPENDENT
PERMEABILITY ..................................................................................................92
6.3.1 Effects of Shape of Pressure Dependent Permeability Curves (Stress
Factor, S) ..........................................................................................................92
6.3.2 Effects of Stimulated Reservoir Volume (SRV) .............................................96
6.3.3 History Matching Using Pressure Dependent Permeability outside SRV .......99
6.3.4 Effect of Native Reservoir Permeability – Well B ........................................103
6.3.5 History Matching Using Pressure Dependent Permeability both Inside and
Outside SRV ..................................................................................................106
6.4 APPROXIMATION OF GEOMECHANICAL EFFECTS IN
CONVENTIONAL RESERVOIR SIMULATOR USING UNIAXIAL
DEFORMATION THEORY ...............................................................................110
6.5 HISTORY MATCHING USING PRESSURE DEPENDENT
PERMEABILITY BOTH INSIDE AND OUTSIDE SRV – COUPLED
PRODUCTION GEOMECHANICAL MODELS ..............................................116
6.6 HISTORY MATCHING USING PRESSURE DEPENDENT
PERMEABILITY BOTH INSIDE AND OUTSIDE SRV – COUPLED
PRODUCTION GEOMECHANICAL MODELS – VARIABLE MEAN
TOTAL STRESS CASES ....................................................................................120
6.7 SUMMARY ...........................................................................................................125

CHAPTER SEVEN: HISTORY MATCHING OF FIELD INJECTION PRESSURE


USING UNCOUPLED AND FULLY COUPLED GEOMECHANICAL
INJECTION MODELS ...........................................................................................127
7.1 MODELING OF FRACTURE PROPAGATION .................................................128
7.2 THEORY OF SIMPLIFIED FRACTURE MODEL .............................................129
7.3 HISTORY MATCHING OF FIELD INJECTION PRESSURE USING
UNCOUPLED INJECTION MODELS ..............................................................131
7.3.1 Effects of Permeability Reduction Factor ( ............................................134
7.3.2 Effects of Limiting Length of Fracture Propagation .....................................138
7.4 HISTORY MATCHING OF FIELD INJECTION PRESSURE USING A
FULLY COUPLED GEOMECHANICAL INJECTION MODEL .....................141
viii
7.4.1 Effects of Limiting Length of Fracture Propagation .....................................144
7.4.2 Effects of Biot’s Constant ( )........................................................................148
7.5 FAILURE PREDICTIONS....................................................................................151
7.5.1 Tensile Failure ...............................................................................................152
7.5.2 Shear Failure ..................................................................................................152
7.5.3 Mohr – Coulomb Shear Failure Criterion .....................................................153
7.6 SUMMARY ...........................................................................................................165

CHAPTER EIGHT: RESULTS AND CONCLUSIONS ................................................166

REFERENCES ................................................................................................................169

APPENDIX A: SELECTED DATA FILES ....................................................................181

ix
List of Tables

Table 3.1 Rock and fluid properties of well A and B (input to reservoir simulator) ........ 36

Table 3.2 Proppant types in well A and B ........................................................................ 39

Table 3.3 Stress and rock properties for geomechanical modeling of well A and B ........ 43

Table 4.1 Baseline permeability of sand used in well A and B ........................................ 54

Table 5.1 Fracture properties of Case 5.1.1 – 2 for well A............................................... 63

Table 5.2 Fracture properties of Case 5.1.1 – 2 for well B ............................................... 63

Table 5.3 Dimensionless fracture conductivity of Cases 5.1.1 - 2 for well A and B........ 69

Table 5.4 Fracture properties of Case 5.2 for well A and B ............................................. 71

Table 5.5 SRVand initial permeability multiplier of Cases 5.3.1 – 6 for well A and B ... 75

Table 5.6 Fracture properties of Case 5.4 for well A and B ............................................. 82

Table 6.1 Fracture properties of Case 5.3.4 and 6.1 for well A ........................................ 88

Table 6.2 Fracture properties of Case 5.3.4 and 6.1 for well B ........................................ 89

Table 6.3 Stress factor (S) of Cases 6.3.1.1 - 4 for well A and B ..................................... 93

Table 6.4 Fracture properties of Case 6.3.2 for well A..................................................... 96

Table 6.5 Fracture properties of Case 6.3.2 for well B ..................................................... 97

Table 6.6 Fracture properties of Case 6.3.3.2 for well A and B ..................................... 101

Table 6.7 Fracture properties of Cases 6.3.4.1 - 2 for well B ......................................... 104

Table 6.8 Fracture properties of Case 6.3.5 both for well A and B ................................ 107

Table 6.9 Fracture properties of Cases 6.6 both for well A and B ................................. 122

Table 7.1 Initial and modified stresses for injection cases – Well A and B ................... 132

Table 7.2 Data for transmissibility calculations in injection cases – Well A and B ....... 133

Table 7.3 Permeability reduction factor of Cases 7.3.1.1-4 both for well A and B........ 135

Table 7.4 Parameters varied in injection Cases 7.3.2.1 – 2 – Well A and B .................. 139

x
Table 7.5 Parameters varied in coupled injection Cases 7.4.1.1 – 4 – Well A ............... 144

Table 7.6 Parameters varied in injection cases 7.4.2.1 – 2 – Well A ............................. 149

Table 7.7 Mohr-Coulomb circles input parameters – Block 1 and 2 – Well A .............. 159

xi
List of Figures and Illustrations

Figure 1.1 Natural gas resource triangle (from Holditch, 2006)......................................... 2

Figure 3.1 Gas compressibility factor and viscosity of well A and B .............................. 35

Figure 3.2 Proppant concentration of well A from field treatment report ........................ 37

Figure 3.3 Proppant concentration of well B from field treatment report ........................ 37

Figure 3.4 Injection rates of well A and B from field treatment report ............................ 38

Figure 3.5 Bottomhole field injection pressure - well A and B ........................................ 38

Figure 3.6 Field gas rate and cumulative production of well A........................................ 40

Figure 3.7 Field gas rate and cumulative production of well B ........................................ 40

Figure 3.8 Horizontal stress trajectories across western Canada sedimentary basin
(from Bell and Babcock, 1986 as reproduced by Mossop and Shetsen (1994)) ....... 44

Figure 4.1 Well completion and fracture spacing – well A and B .................................... 52

Table 4.2 Calculated fracture properties of sands in well A ............................................. 56

Table 4.3 Calculated fracture properties of sands in well B ............................................. 57

Figure 5.1 Flow chart of simulation cases – Chapter 5..................................................... 61

Figure 5.2 Comparison of field data and simulation results of Cases 5.1.1 - 2 - well A .. 66

Figure 5.3 Comparison of field data and simulation results of Cases 5.1.1 - 2 - well B .. 66

Figure 5.4 Dimensionless fracture conductivity of Cases 5.1.1 - 2 – Well A and B ........ 70

Figure 5.5 Comparison of field data and simulation results-Cases 5.1.2 & 5.2 -well A .. 72

Figure 5.6 Comparison of field data and simulation results-Cases 5.1.2 & 5.2 - well B . 72

Figure 5.7 Stimulated reservoir volume (SRV) schematic – Complex fracture


geometry representation ............................................................................................ 75

Figure 5.8 Comparison of field data and simulation results of Cases 5.3.1 - 3 - well A .. 77

Figure 5.9 Comparison of field data and simulation results of Cases 5.3.4 - 6 - well A .. 77

Figure 5.10 Comparison of field data and simulation results of Cases 5.3.1 - 3-well B .. 78

xii
Figure 5.11 Comparison of field data and simulation results of Cases 5.3.4 - 6 -well B . 78

Figure 5.12 Areal view of fracture in x-y plane – Flow pattern representation................ 79

Figure 5.13 Time dependent permeability multipliers-History matched case–Well A .... 82

Figure 5.14 History matching of well A using time dependent permeability ................... 83

Figure 5.15 History matching of well B using time dependent permeability ................... 83

Figure 6.1 Flow chart of simulation cases – Chapter 6..................................................... 87

Figure 6.2 Comparison of field data and simulation results of Case 6.1 - well A ............ 90

Figure 6.3 Comparison of field data and simulation results of Case 6.1 - well B ............ 90

Figure 6.4 Pressure dependent permeability curves of Cases 6.3.1.1 – 4 - well A & B ... 94

Figure 6.5 Comparison of field data and simulation results of Cases 6.3.1.1-4-well A ... 94

Figure 6.6 Comparison of field data and simulation results of Cases 6.3.1.1-4-well B ... 95

Figure 6.7 Comparison of field data and simulation results of Case 6.3.2 - well A ......... 98

Figure 6.8 Comparison of field data and simulation results of Case 6.3.2 - well B ......... 98

Figure 6.9 Comparison of field data and simulation results of Case 6.3.3.1 - well A .... 100

Figure 6.10 Comparison of field data and simulation results of Case 6.3.3.1 - well B .. 100

Figure 6.11 History matched case using PDPM outside SRV - well A.......................... 102

Figure 6.12 History matched case using PDPM outside SRV - well B .......................... 102

Figure 6.13 Comparison of field data and simulation results of Case 6.3.4.1-2-well B . 105

Figure 6.14 History matched case using PDPM outside SRV and modified reservoir
permeability - well B .............................................................................................. 106

Figure 6.15 Pressure and effective stress dependent permeability curves at stress
factor of 2.7 and 3.9 - well A and B........................................................................ 108

Figure 6.16 History matched case using PDPM both inside and outside SRV- well A . 108

Figure 6.17 History matched case using PDPM both inside and outside SRV - well B 109

Figure 6.18 Comparison of field data and simulation results of uncoupled and coupled
cases: Constant mean total stress - well A .............................................................. 117

xiii
Figure 6.19 Comparison of field data and simulation results of uncoupled and coupled
cases: Constant mean total stress - well B .............................................................. 117

Figure 6.20 Coupling of reservoir and geomechanical simulator – An overview .......... 118

Figure 6.21 Plot of mean effective stress and fluid pressure – well A ........................... 119

Figure 6.22 Permeability multipliers and mean effective stress for constant and
variable mean total stress cases – well A ................................................................ 120

Figure 6.23 Pressure and effective stress dependent permeability curves at stress
factor of 2.7, 3.9, 6 and 8.4 - well A ....................................................................... 121

Figure 6.24 Pressure and effective stress dependent permeability curves at stress
factor of 2.7, 3.9, 6 and 8.4 - well B ....................................................................... 122

Figure 6.25 Comparison of field data and simulation results of uncoupled and coupled
cases: Variable mean total stress - well A............................................................... 123

Figure 6.26 Comparison of field data and simulation results of uncoupled and coupled
cases: Variable mean total stress - well B ............................................................... 124

Figure 7.1 Flow chart of simulation cases – Chapter 7................................................... 128

Figure 7.2 Pressure and effective stress dependent permeability curves at stress factor
of 6 (uncoupled injection cases) - well A and B ..................................................... 134

Figure 7.3 Transmissibility multipliers – uncoupled injection cases - well A................ 136

Figure 7.4 Comparison of simulation and field BHIP of Case 7.3.1.1-3- well A........... 137

Figure 7.5 Comparison of simulation and field BHIP of Case 7.3.1.1-3 - well B .......... 137

Figure 7.6 Comparison of simulation and field BHIP of Case 7.3.2.1- well A .............. 139

Figure 7.7 Comparison of simulation and field BHIP of Case 7.3.2.1-2- well B ........... 140

Figure 7.8 Effective stress dependent permeability curve - coupled injection - well A . 143

Figure 7.9 Transmissibility multipliers for coupled injection cases - well A ................. 143

Figure 7.10 Comparison of simulation and field BHIP of Cases 7.4.1.1-4 - well A ...... 145

Figure 7.11 Proppant concentration, BHIP and rate - well A ......................................... 147

Figure 7.12 History matched case – Coupled simulation - well A ................................. 148

Figure 7.13 Comparison of simulation and field BHIP of Cases 7.4.1-4 - well A ......... 149

xiv
Figure 7.14 Minimum effective stress and BHIP at different Biot’s constant numbers
– Coupled cases - well A ........................................................................................ 151

Figure 7.15 Schematics of tensile and shear failure under normal loading .................... 153

Figure 7.16 Mohr-Coulomb failure criterion for critical stress state – A graphical
representation (from Faejer et al., 2008) ................................................................. 154

Figure 7.17 Fluid pressure after 236 days and selected grid block locations – Well A.. 156

Figure 7.18 Minimum effective stress block 1 – 4 – history matched case – Well A .... 157

Figure 7.19 Total in-situ stresses and BHIP for block 1 and 2 –– Well A ..................... 157

Figure 7.20 Mohr – Coulomb failure envelope for Co= 46570psi – Well A .................. 159

Figure 7.21 Stress level in Fracture # 4 after 237mins of injection – Co= 46570 psi –
YZ cross section – Well A ...................................................................................... 160

Figure 7.22 Stress level in Fracture # 4 after 237mins of injection – Co= 46570 psi –
XY areal view – Well A.......................................................................................... 161

Figure 7.23 Mohr – Coulomb failure envelope for Co = 9314 psi – Well A .................. 162

Figure 7.24 Stress level in Fracture # 4 after 237 mins of injection – Co= 9314 psi –
YZ cross section – Well A ...................................................................................... 162

Figure 7.25 Stress level in Fracture # 4 after 237 mins of injection – Co= 9314 psi –
XY areal view – Well A.......................................................................................... 163

Figure 7.26 Mohr – Coulomb failure envelope for Co= 4657psi –– Well A .................. 163

Figure 7.27 Stress level in Fracture # 4 after 141 mins of injection – Co= 4657 psi –
YZ cross section – Well A ...................................................................................... 164

Figure 7.28 Stress level in Fracture # 4 after 141 mins of injection – Co= 4657 psi –
XY areal view – Well A.......................................................................................... 164

xv
List of Symbols, Abbreviations and Nomenclature

cross sectional area

API American Petroleum Institute

fracture area in y-direction

matrix area in y-direction

half aperture distance (fracture half width)

constant defining conductivity / porosity

BHIP bottomhole injection pressure

BHP bottomhole flowing pressure

formation volume factor of phase l

conversion constant

bulk compressibility

fracture conductivity

fracture compressibility at zero stress

grain compressibility

average pore compressibility

CO2 carbon dioxide

pore compressibility

solid grain (matrix) compressibility

constant depending on grain packing

uniaxial compressive strength

xvi
maximum pore compressibility

d average fracture spacing

elastic modulus

dimensionless fracture conductivity

flow factor

shear modulus or modulus of rigidity

acceleration of gravity

poroelastic constant

formation net pay thickness or pay zone

fracture height

root mean square value of height distribution

IPM initial permeability multiplier

formation permeability

gas apparent permeability

bulk modulus of grain

fracture permeability

Kh horizontal formation permeability

effective fluid permeability

matrix permeability

microcracks permeability

dimensionless parameter that describes

conductivity for fluid seepage in plugged pores

xvii
relative permeability of phase l

solid grain modulus

baseline permeability of sand

Kv vertical formation permeability

specific permeability to water

pressure dependent formation permeability

permeability at initial pressure or effective stress

Klinkenberg permeability (no gas slippage)

permeability at 1000 net confining pressure

crack length per unit area

fracture half length

mass of sand in formation

number of thickness

net overburden pressure

net confining pressure

PDPM pressure dependent permeability multipliers

overburden pressure

external pressure

fluid pressure

initial fluid pressure

fracture opening or closing pressure

fluid pressure in fracture at a distance x

xviii
internal pore or fluid pressure

equivalent pressure due to cementation and permanent

deformation of grains

confining pressure

maximum net overburden pressure

minimum net overburden pressure

PVT pressure volume and temperature

fluid pressure at the wellbore

effective elastic modulus

arithmetic mean of gas pressure in core

flow rate of phase l

local fluid flow rate

constant, permeability reduction factor

radius of curvature

wellbore radius

effective wellbore radius

stress factor

skin

combined standard deviation of peak heights

SRV stimulated reservoir volume

shear strength or Cohesion

transmissibility

Tcf trillion cubic feet

xix
transmissibility of fracture in y-direction

matrix transmissibility in y-direction

transmissibility multiplier in y-direction

transmissibility multiplier in z-direction

transmissibility multiplier

ultimate tensile strength or tensile strength

t time

volume of sand (downhole volume of sand)

X-link cross link

downhole volume

fracture volume

volume of total fracturing fluid injected

volume of fluid leaked off from fracture face into formation

grid width

fracture width

total fracture length in y-direction (y = 2Lf)

shear strain

Biot’s constant

constant from curve fitting experimental data

failure plane orientation angle

empirical coefficient

strain in corresponding direction

xx
volumetric strain

poroelastic constant

l phases i.e., gas, oil, and water

lame’s 1st parameter

fracture fluid viscosity

viscosity of fluid of phase l

formation bulk density of the overburden

density of sand

external stress

maximum horizontal total stress

minimum horizontal total stress

initial minimum horizontal stress

minimum horizontal stress at corresponding time

stress in corresponding direction

total mean stress

initial total mean stress

vertical total stress

effective stress

minimum horizontal effective stress

mean effective stress

maximum effective stress

minimum effective stress

xxi
effective stress in x-direction

effective stress in y-direction

or vertical effective stress

initial effective stress

initial effective stress in x-direction

initial effective stress in y-direction

tangential effective stress

shear Stress

Poisson’s ration

friction angle

formation porosity

pressure dependent formation porosity

sand porosity

instantaneous porosity

porosity at initial pressure or effective stress

difference operator

pressure drop

grid size in y-direction

increase in effective stress

Subscripts

i x, y, and z - direction

ij xy, xz, and yz – direction

xxii
CHAPTER ONE: INTRODUCTION

1.1 BACKGROUND

Holditch (2006) defined the tight gas reservoirs as

“Reservoirs that cannot be produced at economic flow rates nor recover economic

volumes of natural gas unless the well is stimulated by a large hydraulic fracture

treatment, by a horizontal wellbore, or by use of multilateral wellbores.”

The importance of unconventional reservoirs increased with advancements in horizontal

well drilling and hydraulic fracturing technology over the past decade. Tight gas

reservoirs produce at low rate over longer period of time due to low permeability.

Compared with conventional reservoir much closer well spacing is needed for tight gas

reservoir to get economic recovery. The concept of resource triangle used by Masters

(1979) gives us very good idea about importance of unconventional resources. Figure 1

illustrates that high quality and highly permeable hydrocarbons, which occur in small

volumes, lay on top of triangle while abundant hydrocarbons having very small

permeability lay at the bottom of triangle. The inadequate gas price and existing

technology make it difficult to recover unconventional hydrocarbons at the base of the

triangle at full scale. The ease to recover the hydrocarbons with current technology

decreases downward through the triangle. The history of tight gas sand exploitation traces

back to 1950’s and 1960’s and the development accelerated with the advent of hydraulic

fracturing technology and developments in multilateral well completions. Increased gas

1
prices and improved technology has led to development of several tight gas sand plays all

over the world, most notably in North America.

Limited knowledge of geology and engineering of unconventional gas resources, natural

gas policies and unfavourable market conditions and shortage of expertise currently

obstructs the exploitation and development of tight gas reservoirs world-wide except in

North America. 58 Tcf of gas has been produced from tight gas reservoirs in U. S.

throughout the year 2000 and it is estimated that 20% of gas production in U.S. is

currently from tight gas reservoirs as reported by Holditch (2006). According to Rogner

(1996) as reported by Kawata and Fujita (2001) there are about 7405 Tcf of gas in place

in tight gas sandstones all over the world.

Figure 1.1 Natural gas resource triangle (from Holditch, 2006)

Rapid growth of global energy consumption, increasing demand for fuel and faster

decline of conventional energy resources shift the focus of global energy producers to

2
development of unconventional resources. To fully exploit the unconventional resources

attention must be paid to understanding of the geology and engineering. New methods

and techniques should be developed based on better understanding of the theory to

maximize recovery.

1.2 RESEARCH OBJECTIVES

The current project is part of a research consortium which studies new simulation

methods for the development and exploitation of unconventional natural resources. The

emphasis of the work is on better understanding and modeling of reservoir engineering

and geomechanical aspects of stimulation and production of tight gas and shale

reservoirs. The research is divided into two main sections.

1.2.1 Production Modeling of Multifractured Horizontal Wells

The main purpose of this project is to study different fracturing techniques used in

hydraulic fracturing, and the effect of fracturing and geomechanical behavior of the

reservoir on well productivity. The study utilized stimulation and production data for two

horizontal wells in a tight gas play. The objectives of this study are discussed below:

 Model production performance of both wells with different fracture properties.

 Estimate fracture parameters using downhole injected proppant volume.

 Investigate effects of fracture parameters (fracture length, conductivity and

width), spacing, fracture density and assumed size of Stimulated Reservoir

Volume (SRV) on gas rate and gas production.

3
 Carry out preliminary history matching for both wells with available data from the

field using manual change of reservoir permeability with time.

 Investigate the effects of pressure / stress dependent permeability on results, both

inside and outside stimulated reservoir volume.

 Carry out history matching using pressure / stress dependent permeability using

both uncoupled and fully coupled geomechanical models.

 Approximate the poroelastic effects (i.e., geomechanical effects) in a conventional

reservoir simulator without running fully coupled geomechanical simulator.

Apache Canada, which participates in the consortium, provided field data for the purpose

of this study. They provided data for 2 different wells, which will be referred to here as

Well A and Well B. The wells were drilled in the same formation but at different

locations. As both wells were completed in tight gas sand reservoir, they were stimulated

by hydraulic fracturing to get maximum recovery.

Two different kinds of fracturing techniques were used in each well, and they will be

represented by following names.

 X-link Gelled Water Frac with CO2 (Hybrid)-----------------------Well A

 Slick Water Frac (Ungelled Water Frac) with CO2------------------Well B

1.2.2 Injection Modeling of Multifractured Horizontal Wells

In this part of project the objective is to study the injection (stimulation) process itself.

The main tasks are:

4
 To model and history match field injection bottom hole pressure using both

uncoupled and fully coupled geomechanical models.

 To study in detail the effects of Biot’s constant and permeability reduction factor,

as history matching parameters, on field injection pressure.

 Carry out a sensitivity study of these parameters on injection pressure.

 Investigate the shear failure mechanisms and the extent of the Stimulated

Reservoir Volume (SRV) as a function of rock mechanics data.

1.3 THESIS ORGANIZATION

The major tasks, which have been completed in the course of this study, are described in

Chapters organized as follows:

1. Literature Review is summarized in Chapter 2.

2. Chapter 3 gives brief summary of field data that includes basic data required for

reservoir study, stimulation data from field, field production data and

geomechanical data.

3. Chapter 4 discusses modeling of static fractures using field stimulation data,

estimation of fracture parameters and representation of these fractures in

simulator.

4. Fracture conductivity can significantly affects the production results. The effects

of fracture conductivity along with other parameters such as net pay thickness and

5
stimulated reservoir volume on gas rate cumulative production are discussed in

Chapter 5.

5. In Chapter 6, the effects of pressure and stress dependent permeability and shape

of permeability curves on production are discussed. Use of pressure and stress

dependent permeability both inside and outside stimulated reservoir volume is

essential for history matching, and is implemented using both uncoupled and fully

coupled geomechanical models. A new method based on uniaxial deformation

theory is proposed to approximate poroelastic effects and model geomechanical

effects in conventional reservoir simulator (without running fully coupled

geomechanical model).

6. History matching of field injection pressure with simulation results using

uncoupled and fully coupled injection model is given in Chapter 7. Simplified

fracture models, fracture propagation modeling, effects of permeability reduction

factor, and Biot’s constant are also discussed in great detail in Chapter 7.

7. Summary of results and conclusions is discussed in Chapter 8.

In addition, research was carried out on two related topics. These are not included here

partially because of space limitations, and also because they require some additional

work. However, they are briefly described below. We intend to publish this work

separately.

 First, fracture density and location of fractures (fracture placement) plays a very

important role in production after stimulation treatment. It is often the case that

some fractures are not contributing to production for various reasons. Cases

6
having different sets of fractures of several combinations were simulated and

categorized in classes based on equal effects on gas rate and cumulative

production.

 Second, the flow of gas in shale is more complex compared to conventional

reservoirs, and additional physics such as slippage desorption, and diffusion may

play a large role. A study of these phenomena on a pore scale has been carried out

jointly with other researches, and shows that the effects can be significant if

permeability is small. The effects are more dominant and considerable in shale

gas reservoirs but initial investigation revealed that these factors are equally

important for tight gas sands.

Finally, the use of units must be discussed. In this work, we use the field (Imperial)

system of units unless stated otherwise. The reason is that the large amount of data we are

dealing with is in these units which are customary in the petroleum business.

7
CHAPTER TWO: LITERATURE REVIEW

The process of creating a single fracture or a fracture system in porous medium by

injecting fluid at very high pressure is termed hydraulic fracturing (well stimulation). In

the process, fluid pressure must overcome the compressive stresses and tensile strength of

the rock, which defines the fracturing pressure during the fracturing job. In a typical

hydraulic fracturing process, two wings of a crack (fracture) are created initially by

injecting fracturing fluid at very high rate and pressure. The clean fluid which is called

pad is injected first in order to extend crack growth deep into the formation and create a

sufficient width for slurry injection. Specially designed slurry containing proppants of

different types is pumped following the pad. The fracturing fluid may contain borate,

guar gum, x-linked gels, alcohols and various other fluids in different combinations to

carry proppant and further extend the fracture and create a highly conductive path.

Chemicals used during fracturing degrade at reservoir conditions which help to recover

injected fluid during flow back.

Klepper No. 1 was the first well to be hydraulically fractured in Hugoton gas field in

western Kansas in 1947. The concept of hydraulic fracturing has evolved from already

existing practices in petroleum industry at that time such as water injection, acidizing of

oil and gas wells and cementing. The process was adopted very rapidly after its initial

successful jobs and performances so that more than 10,000 jobs had been performed by

the end of 1955. Clark (1949) was the first person to introduce hydraulic fracturing to

petroleum industry by name of “Hydrafrac”. Hubbert and Willis (1957) established that

8
the fracture orientation is perpendicular to the direction of the least principal compressive

stress, which in turn can be assumed from the faulting of a region. The minimum

principal stress is usually horizontal in tectonically relaxed areas described by normal

faulting and fractures produced in such areas should be vertical. The minimum principal

stress may be vertical in tectonically active area described by folding or thrust faulting

and fractures produced should be horizontal. They were the first to point out that

fractures produced should be perpendicular to least principal stress irrespective of fluid

type and supported their argument by experiment. Harrison et al. (1953) strongly favored

theory presented by Hubbert and Willis.

In the beginning the process was used as a recovery process in most of the wells.

Nowadays hydraulic fracturing is an essential part of well completion and plays a very

important role in exploitation and development of unconventional reservoirs such as tight

gas sands, shale gas, shale oil and heavy oil. Recently it is being used in applications such

as waterfracs, waterflood-induced fracturing, produced water re-injection, drilling waste

disposal and induced fracturing in thermal recovery methods.

2.1 PRESSURE / STRESS DEPENDENT ROCK PROPERTIES

2.1.1 Experimental Work

The non-rigid and deformable nature of porous media and its effect on flow properties

was recognised as early as 1928 (Meinzer, 1928; Jacob, 1940, 1950). Earlier laboratory

data on reservoir rocks properties were obtained by ignoring overburden pressure and this

was continued until the importance of pressure / stress dependence of rock properties was

9
recognized. Carpenter and Spencer (1940) conducted experiments on oil-bearing

consolidated sandstones and showed that change in porosity with overburden pressure is

small. Fatt and Davis (1952) in their experiments on eight typical consolidated oil bearing

sandstones showed permeability changes with overburden pressure. They concluded that

most of the decrease takes place from zero to 3000 psi range of overburden pressure and

reduction in permeability generally increases as initial permeability increases. The effect

of overburden pressure on pore volume compressibility and relative permeability is small

and can usually be ignored but permeability dependence on overburden pressure is very

strong and reduction in permeability is larger for low permeability rocks (Fatt, 1953).

Hall (1953) conducted experiments on samples from an under-saturated reservoir and

established that total or effective compressibility of any reservoir rock is the results of

two separate factors, i.e., an expansion of the individual rock grains as the surrounding

fluid pressure decreases and the additional formation compaction because of increase in

effective overburden pressure due to pressure decline. McLatchie et al. (1958) and Wyble

(1958) performed experiments on various oil-bearing sandstones over a range of effective

burden pressure of 0 – 8000psi and 0 – 5000psi respectively. They concluded that at

certain given effective overburden pressure the percentage reduction in permeability is

larger for low permeability rocks. Anisotropy in stress is also stress-dependent and

permeability reduction under uniform hydrostatic loading in laboratory is always greater

than that under non uniform loading, i.e., permeability reduction is always overestimated

or reported permeability is always underestimated (Gray and Fatt, 1963). Vairogs et al.

(1971) performed experiments on 10 core samples of initial permeability ranging from

10
0.04 mD to 199 mD over wide range of net confining pressure up to 20,000 psi. They

made the following observations from experiments:

 The effect of stress on tight rock is far greater than for highly permeable rocks

and the reason being is smaller pore radii in tighter formation. The decrease in

flow capacity due to increasing confining pressure for smaller pore radii is much

larger than in larger pores.

 The presence of fractures and shale streaks further accelerate permeability

reduction due to increasing confining pressure.

Thomas and Ward (1972) performed experiments on a variety of tight sandstone core

samples. Significant decrease in permeability with increasing net burden pressure was

observed and most of the decrease took place at pressure to 3000 psi. According to them

relative permeability and porosity does not significantly changes with net overburden

pressure while permeability decreases tremendously with increasing net overburden

pressure in presence of fractures. Brighenti (1989) studied core samples of permeability

range of 0.0009 – 200 mD and different mineralogical characteristics in the laboratory.

He found correlations between normalized permeability and applied stress (net confining

pressure) from experimental data and results from those correlations were in good

agreement with Jones and Owens (1980) and Ostensen (1983) model for stress-dependent

permeability of tight gas sandstones. He reported that reduction in permeability is larger

for low permeability core samples and highlighted that the effect is more pronounced in

presence of fractures.

11
Natural fractures or fissures are considered to be the main source of hydrocarbon

transport in low permeability reservoirs from super low permeability matrix blocks

(Warpinski, 1991; Buchsteiner et al., 1993; Lorenz, 1999). Matrix rock permeability is

typically < 1 μD under in-situ conditions for low permeability reservoirs. Fissures or

natural fractures exist in many of these formations and their presence is critical to the

recovery of hydrocarbon from these reservoirs (Warpinski, 1991). Permeability of these

reservoirs increases one to two orders of magnitude during injection due to decrease of

effective stress, and decreases during drawdown or production due to increase of

effective stress. Reyes and Osisanya (2002) performed experiments on core samples of

four shale types (Wapanuka, Wilcox, Atoka and Catoosa formations in Oklahoma) for

studying effects of in-situ stresses on porosity and permeability. They developed

empirical correlations from experimental data and found that porosity and permeability

are both function of effective stress and must be considered during well development

programs from very beginning. Lei et al. (2007) performed experiments on five core

samples over the range of 2 – 25 MPa effective stress. According to them sensitivity of

permeability to variations in stress is due to the significant change in the bearing

skeletons, solid particles, and pore throats. Pores with oval or polygonal shapes are more

resistant to deformation while pores with flat shapes or silt like structure are more stress

sensitive. Permeability is mainly controlled by pore throats of radii less than 1 μm in tight

formations.

12
2.1.2 Mathematical Correlations

Dobrynin (1962) developed mathematical equations for physical properties such as

porosity, density, permeability, resistance and velocity of elastic waves and verified the

results with experimental data from consolidated sandstones over the range of 0 – 20,000

psi net overburden pressure. The changes of these physical rock properties were due to

changes in pore space of the rocks. Pore compressibility and permeability expressions as

function of net over burden pressure were given by

where = Porosity

= Permeability

= Maximum pore compressibility

= Net overburden pressure =

= Overburden pressure

= Internal pore pressure or fluid pressure

= Empirical coefficient

= Pore compressibility

13
Maximum pore compressibility can be determined by extrapolation of the experimental

curves of pore compressibility to zero pressure using Cartesian co-ordinates. The

empirical coefficient  depends on pore size distribution and pore compressibility; it can

be found by solving Equation (2.4).

Gangi (1978) theoretically derived mathematical equations for permeability and porosity

variation with pressure in porous medium using Hertzian theory of deformation of

spheres and his result is given as;

where = Constant depending on grain packing

= Permeability at initial pressure

= Porosity at initial pressure

= External pressure =

= Radius of spherical grains

= Equivalent pressure due to cementation and permanent deformation of

grains

= Effective elastic modulus of the grains

= Bulk modulus of the grains

14
= Poisson’s ratio of the grains

The theoretical curves generated using the above correlations are found to be in good

agreement with experimental data but this expression needs information such as

cementation of grains, Poisson’s ratio of rock, bulk grain modulus and approximation of

the constant for grain packing.

Jones and Owens (1980) conducted compressibility and flow tests on more than 100 tight

gas sand core samples. They found the following correlations for permeability, gas

slippage and water permeability:

where = Stress factor

= Confining pressure

= Gas apparent permeability

= Klinkenberg permeability (no gas slippage)

= Arithmetic mean of gas pressure in core during flow of gas

= Specific permeability to water

According to them flow in tight gas sandstones is mainly controlled by slit like micro

cracks rather than round shape capillary tubes and they showed that the cubic root of

permeability varies linearly with log of confining stress. They pointed out that routine

permeability measurement (at ambient pressure) should be corrected to 1,000 psi

15
confining pressure, which is the main reason for the appearance of 1,000 in the above

equation. Rocks having lower permeability are affected more by net effective stress than

high permeable rocks. Pore volume compressibility (porosity change with pressure) is

only very slightly affected by increasing confining pressure and hence this nonlinearity

can be ignored.

Jennings et al. (1981) proposed a model to predict permeability and electrical

conductivity of formations. The model assumes that rock composed of irregular and

small pores which are interconnected by thin cracks and microfractures can be modeled

by rectangular slits rather than by round shape capillary tubes. In their model

permeability is related to confining pressure by third order polynomials.

where = Coefficients of cubic equation relating permeability to net

confining pressure

= Net confining pressure

The A coefficients in the above equation can be found by curve fitting of polynomial to

experimental data.

Walsh (1981) developed an alternative asperity-controlled model for fracture

permeability vs. effective stress. This model is based on experimental data on the shape

of surface roughness. By approximating the Gaussian asperity height distribution with an

exponential, Walsh theoretically derived Jones and Owens (1980) correlation. He

confirmed results from his theoretical model given below by Equation (2.12) with

available experimental data from literature.

16
where = Root mean square value of height distribution

= Permeability at initial effective stress

= Half aperture distance or half fracture width

= Effective stress

= Initial effective stress

The first term in Equation (2.12) is the effect of aperture and second term is the effect of

tortuosity. Walsh suggested that based on experimental results the second term in the

above equation can be neglected except for very high effective stress. The analysis

suggested that K1/3 should be linearly related to .

Ostensen (1983) established the existence of network of grain-boundary cracks in tight

gas sandstones. He developed stress dependent permeability model based on flow

through micro cracks in which crack width is controlled by elastic deformation of surface

asperities.

Where = Microcrack permeability

= Crack length per unit area

= Combined standard deviation of peak heights

= Elastic modulus

17
= Radius of curvature

= Poisson’s ratio

= External stress

His model strongly supports the theory that the permeability in tight gas sandstones is

controlled by flow through micro cracks. He proved (using available data from tight gas

sandstones) that square root of permeability varies linearly with log of stress and his

correlation is better than the one presented by Jones and Owens (1980). He was of the

opinion that if micro cracks controls permeability in tight gas sandstones, then the same

could be true for pore compressibility (Porosity). Plot of total porosity with log of

confining pressure should give straight line, which was confirmed with available data.

McKee et al. (1988) have derived relationship for permeability, porosity, and density as

functions of effective stress by using Carman-Kozeny equation. The major assumptions

are incompressible solid matrix and volume changes resulting from compression were

ascribed to pore spaces. Eight core samples, i.e., four coals, one granite, two sandstones,

and one clay, were used to verify applicability of correlations and excellent matches were

found. The resulting relations are:

where = Average pore compressibility

= Increase in effective stress = =

= Pressure drop =

18
= Permeability at initial effective stress

= Porosity at initial effective stress

and s is a constant for linearly elastic materials and rocks (Walsh, 1981).

Buchsteiner et al. (1993) developed equations for fracture permeability and permeability

anisotropy vs. effective stress for fissured or natural fractures dominant reservoirs, which

are stress sensitive. An increase in effective stress due to initial large drawdown results in

high initial rates termed “flush” production provided by rapid drainage of fracture and

fracture network. Flush production substantially reduces permeability, thus inducing

permeability anisotropy and sometimes altering direction of maximum and minimum

permeabilities as a result of depletion.

where = Conductivity / Porosity

Liu and Civan (1995) developed correlation both for permeability and porosity using

Carman-Kozeny equation and the pore compressibility concept, which is given as;

where = Fracture compressibility at zero stress

= Constant determined by curve fitting experimental data

= Grain compressibility

19
= instantaneous porosity during particulate processes

= Dimensionless parameter that describes conductivity for fluid seepage in

plugged pores

= Flow factor, it is defined as the fraction unplugged pore throats

As mentioned earlier, Reyes and Osisanya (2000) developed empirical correlations for

porosity and permeability from experimental dataset of four shale core samples. Their

correlations are in exponential form and depend on the average pore compressibility of

the system. Lei et al. (2007) correlated permeability and effective stress with quadratic

polynomial relation from experimental results on five core samples.

2.2 EFFECTS OF PRESSURE / STRESS DEPENDENT PERMEABILITY ON

GAS PRODUCTION

Vairogs et al. (1971) were the first to study the effects of stress dependent permeability

on gas production. They confirm the conclusions made by preceding authors through

laboratory experiments. They developed a reservoir simulator that incorporated stress

effects for single phase isothermal gas flow to well in a circular reservoir. They assumed

that reduction in permeability caused by changes in radial stresses is small as compared

with that due to the tangential and vertical stresses, and the change in permeability is

therefore a function of average effective (confining) stress defined as

where = Mean effective stress

= Tangential effective stress

20
= Vertical effective stress

They showed that significant decrease in gas production rate was modeled due to

reduction in permeability as gas pressure declines. The authors also concluded that

increasing pressure drawdown will not help to obtain larger recovery in tight gas

reservoirs. Vairogs and Rhoades (1973) have done pressure transient analysis on gas

wells. They showed that drawdown tests in stress sensitive formations are not reliable.

Permeability obtained from semilog analysis for drawdown tests is not consistent and

build up tests give reasonably good results both for early time and late time.

Ostensen (1986) in his paper emphasized that the effect of effective stress on

permeability in tight gas reservoir cannot be ignored and permeability data obtained

under uniformly stressed conditions in laboratory is not actual representation of in-situ

conditions because horizontal stress is usually lower than vertical. He concluded that the

permeability reduction is overestimated or the reported permeability is underestimated by

factor of two or more depending upon cracks directions if measured under uniform

hydrostatic conditions. His finding agrees with the conclusions of Gray and Fatt (1963).

Stress-dependent permeability decreases initial steady state production by 30% for typical

tight gas sand reservoir. Average decrease in stress dependent permeability mainly

controls the late time steady state production.

Warpinski (1991) concluded that low permeability or tight formations require special

care to recover hydrocarbons. Problems that can significantly reduce production from

these fissured formations are leak off, damage, and complex fracturing. Leakoff during

stimulation may be with constant or pressure sensitive permeability and becomes more

severe when natural fractures begin to dilate and results in high leakoff into formation.

21
Liquid damage is transient and will eventually recover but it may take weeks or even

months. Mechanical damage is due to shear slippage and it is permanent. Complex

fracturing will cause high treatment pressure and generally reduce effective half length.

Special care must be taken when formations are initially perforated and fractures are

created otherwise anticipated production may not be achieved.

Significant reduction in production rates in stress sensitive reservoirs is due to rapid

decrease in near-wellbore permeability because of increase in effective stress as a result

of pressure depletion. Thus, reducing bottomhole pressure to increase recovery may not

help because the high pressure drawdown near wellbore will eventually decreases

permeability. The best way to increase production in tight formations is to maintain high

reservoir pressure (Buchsteiner, 1993). Lorenz (1999) reported that although large

fractures may dominate flow at well bore, they must have support from densely populated

smaller and more stress sensitive fractures. Conductivity for fractures width less than 10

μm drops more rapidly than cubic of aperture.

Davies and Davies (2001) indicated in their study that reservoirs characterized by stress-

dependent permeability are dynamic permeability reservoir models. Stress dependent

permeability, both in unconsolidated and consolidated reservoirs, is controlled mainly by

pore geometry, and the rate of change of permeability with stress is a direct function of

rock type. According to their conclusions sand with large pores and initial permeability is

more stress sensitive for unconsolidated sandstones. In consolidated sandstones, sands

with long narrow slit like pores are more stress sensitive than those with long capillary

tube like pores. They showed that stress dependent permeability has significant effects on

the performance of both well and reservoir as a whole.

22
Lei et al. (2007) performed production analysis on low permeability core samples and

concluded that near wellbore permeability decreases up to 30%, which results in

production rate decrease of up to 21%. They emphasized that optimum bottomhole

flowing pressure should be determined from reservoir conditions and rock stress

sensitivity to get maximum recovery. Pressure / stress dependent permeability is equally

important in injection modeling such as for water flooding, water reinjection, etc. Wu et

al. (20008) showed that assuming constant permeability in stress sensitive reservoirs

overestimates pressure response at the wellbore as it ignores the effect of fracture

expansion on fracture permeability. They concluded that for high injection pressure

operation neglecting permeability dependence on pressure will lead to large errors.

2.3 CONVENTIONAL FRACTURE MODELS

Knowledge of fracture evolution and creation (i.e., fracture length, width and height) and

how these properties affect the fracture performance is critical to fracture modeling and

simulation. Fracture propagation in fracture simulation software packages, currently

available in the market, is estimated by solving the material balance of injected fluid. The

total amount of fracturing fluid injected minus the volume of fluid leaked off from the

fracture faces into the formation is the amount of fracturing fluid remaining in fracture

that actually constitutes fracture volume. The fracture volume from material balance of

injected fracturing fluid is written as:

where = Fracture volume

23
= Volume of total fracturing fluid injected

= Volume of fluid leaked off from fracture face into formation

Fracture volume is calculated from Equation (2.21) at any time, which is then used to

estimate fracture parameters using different assumptions about fracture geometry.

2.3.1 Two – Dimensional (2-D) Fracture Models

Two dimensional fracture models only solve for two fractures dimensions, i.e., fracture

width and length (or radius) while fracture height is assumed constant (generally equal to

net pay thickness or pay zone thickness, or determined by stress contrast). Two

dimensional models such as PKN and KGD model are discussed here.

2.3.1.1 PKN Fracture Model

PKN fracture model was first proposed by Perkins and Kern (1961), and was later

improved by Nordgren (1972). PKN fractures model takes the following assumptions to

derive relations for fracture length and width.

 They considered vertically limited fracture, i.e., constant fracture height usually

equal to net pay thickness.

 Fracture in each vertical cross section deforms independently in vertical planes

perpendicular to the main fracture planes, i.e., plane strain conditions exist.

 Fracture vertical cross section is elliptical and maximum fracture width is at the

center of a cross section (ellipse).

24
 The fluid pressure in the fracture is constant in each vertical cross section

(perpendicular to fracture propagation). Fluid pressure is then only a function of

distance from the well, and at the fracture tip is equal to total earth stress

perpendicular to plane of the fracture, i.e., minimum horizontal stress.

Perkins and Kern (1961) derived the equations for fracture length and width vs. time both

for Newtonian and Non Newtonian fluid in laminar and turbulent flow regimes

neglecting fluid loss and rate of change of fracture volume. Nordgren (1972) in his work

later concluded that length at early times must be determined by their formula not

Carter’s formula while width predicted by Perkins and Kern is good approximation to his

solution. Large time fracture length formula is identical to Carter’s formula and width

formula gives superior results to the Perkins and Kern width formula.

PKN fracture width at any location x along the fracture length at a time t is given as

where = Fracture height

= Poisson’s ratio

(x) = Fluid pressure in fracture at a distance x from the well

= Minimum principal horizontal stress

= Shear modulus or Modulus of rigidity

Pressure gradient for restricted, vertical, and elliptical fracture from injection of

Newtonian fluid in laminar flow is

25
Where = Local fluid pressure

= Local fluid flow rate

= fracturing fluid viscosity

= Local fracture width

= coordinate along fracture length

2.3.1.2 KGD Fracture Model

KGD model solves the 2-D vertical crack problem in a horizontal plane. Khristianovic

and Zheltov (1955) were first to investigate the fracture width assuming constant width in

each vertical crossection which was later improved by Geertsma and de Klerk (1969).

KGD model assumes plane strain condition in a horizontal plane, constant fracture height

and fracture tip which closes smoothly.

The maximum fracture width for bi-wing linearly propagating fracture at wellbore (x=0)

at any time t is

Where = Fracture half length

= Fluid pressure at the wellbore (i.e., the bottomhole pressure)

Khristianovic and Zheltov (1955) expressed pressure gradient for laminar fracture fluid

flow and linearly propagating fracture by

26
A very small region is assumed at the very tip of fracture, where there is no fluid flow

and there is infinite flow resistance in that region. Most of the pressure drop occurs at the

tip of the fracture and therefore pressure and flow rate are constant in majority of the

fracture except for small region near fracture tip (Khristianovic and Zheltov, 1955).

Carter’s model is not used presently for fracture propagation modeling but Carter’s leak

off model is still used in fracture modeling. Although PKN and KGD models (2-D

models) have been replaced by fully 3-D and pseudo 3-D models in fracture modeling

these days, they are still available in many fracture modeling software packages as these

models serve as the basis for current 3-D models discussed next.

2.3.2 Three – Dimensional (3-D) Fracture Models

Real fractures are three-dimensional, which means that fracture height, width and length

all grow with respect to time and space. 3-dimensional models allow fracture height to

change at any location along the fracture length and correspondingly the width and its

vertical profile also change throughout the injection period.

In fully three – dimensional fracture models (Clifton and Abou-Sayed, 1981; Cleary,

1983) fracture develops vertically in plane perpendicular to minimum in – situ stress

assuming that the medium through which it can propagate, i.e., formation and the upper

and lower shale, is infinite. The fracture shape and width is obtained from a 3-

dimensional solution of frac opening equations on a half-space. Pseudo-three-

dimensional fracture models (P3D) usually assume that fracture length is relatively large

compared to fracture height so that all cross-sections deform independently of each other

in the lateral direction, and the propagation of the vertical tips can be obtained by a series
27
of 2-D solutions. Lumped 3-D fracture models are based on further simplification of the

P3D fracture models. These models assume that the fracture height growth is governed

by KGD fracture model and the fracture length growth is governed by PKN fracture

model for fractures in well-contained reservoir where fracture length is larger than

fracture height. The corresponding lumped fracture model is called a Lumped Lateral

Model. Analogously, for fractures in poorly-contained reservoir where fracture height is

larger than fracture length, one can assume that the height growth is governed by PKN

fracture model and the fracture length growth is governed by GDK model, resulting in a

Lumped Vertical Model.

All of the fracturing models discussed above are primarily concerned with the created

fracture geometry. However, for production modeling the important part of the solution is

the propped fracture geometry, which must be obtained by simultaneous solution of the

proppant transport during the fracturing job. The propped length and height of the

fracture can be significantly smaller than the created fracture dimensions, and it is a

function of not only the created fracture geometry, but a number of other parameters. In

our work, we will be primarily concerned with the propped fracture dimensions and its

conductivity.

2.4 UNCONVENTIONAL FRACTURE MODELS

The presence of microcracks and natural fractures in tight formation has been known for

some time, but the advancements in hydraulic fracturing technology and interdisciplinary

research in this area helped to understand the complex nature of these reservoirs.

28
Interactions of natural fractures or fissures with single planar fracture form a complex

fracture networks with increased permeability around the main fracture in naturally

fractured reservoir such as tight sands and shales. This area of increased permeability is

often referred to as the “Stimulated Reservoir Volume” (SRV) and this concept is largely

based on the understanding and analysis of microseismic events (Olson et al., 2012). The

branching of hydraulic fractures due to crossing, bypassing, intersecting, arresting due to

shear slippage or dilating, and diversion along natural fractures or fissures give rise to

development of a complex fracture network, modeling of which is referred to as an

Unconventional Fracture Model (UFM). Interactions of natural fractures with the

hydraulic fractures and their contributions to production in unconventional reservoirs are

still a current topic of debate and discussion. Laboratory experiments are a way to

investigate the physics of complex hydraulic fracturing but one potential problem is the

scaling of laboratory results to real field operations (Olson et al., 2012).

Sampath and Keighin (1982) made pore casts from tight gas sand cores. They reported

that thin film inter-granular pores (less than 3 μm across) often observed in sandstones

and tight sands have large number of pores smaller than 1 μm. Walls (1982) performed

experimental study on 10 core samples from Spirit River Formation in Alberta and 2

samples from Cotton Valley Formation in Texas. He reported that pore structure in tight

sands is very narrow and slit like aperture between pores acts like thin cracks, which

provides major connectivity and allow fluids through it. Brower and Morrow (1985)

produced pore casts tight sands and concluded that flow through tight gas sands is

controlled by network of interconnected polyhedral sheet pores, which are somewhat like

the surface of a randomized honeycomb.

29
Warpinski and Teufel (1987) have provided direct evidence of complex fracture

geometry from mineback studies and illustrated the complex nature of actual fractures

created in naturally fractured or fracture dominated reservoir. They reported that geologic

discontinuities – faults, joints, bedding planes, and stress anisotropy have significant

effects on fracture height, length, leakoff, treatment pressure, and proppant transport. The

effects of these discontinuities on hydraulic fracture strongly depend on treatment

pressure, in-situ stresses, orientations of fissures, and permeability. A hydraulic fracture

model that simulates a complex fracture network, referred to as Unconventional Fracture

Model (UFM), had been recently been developed assuming presence of natural fractures

in formation (Weng et al., 2011). They concluded that stress anisotropy, natural fractures,

and interfacial friction play an important in creating fracture network and their results can

be used to improve fracture stimulation design.

Lin and Zhu (2012) presented a semi-analytical model using a source and sink method to

estimate well performance in complex fracture geometry by assigning fracture network of

natural fractures to multiple hydraulic fractures. Each fracture (either hydraulic or

natural) is considered as a series of line sources and they analytically solved the problem

using principle of superposition. Olson et al. (2012) carried out experiments in the

laboratory to examine the effects of cemented natural fractures on hydraulic fracture

propagation. They showed that oblique embedded fractures more likely divert the

hydraulic fracture as compared to orthogonal fractures. They demonstrated in their

experiments that interactions of hydraulic and natural fractures are due to hydraulic

fracture bypassing natural fracture by propagating around it, hydraulic fracture arresting

into a natural fracture and diverting along it, and combination of both.

30
Joint deformation, magnitude and orientation of original and altered stresses, and natural

factures spacing are sources of fracture complexity and interconnectivity (Urbancic,

2012). Wu et al. (2012) presented a method for computing so called “stress shadow”

which may be present during stimulation in multifractured horizontal wells. Stress

shadow is an effect of in-situ stress increase seen when performing a given stage of the

treatment caused by additional stresses created by previous stimulation in the adjacent

treatment stages in a complex fracture network. Stress shadow can significantly affect

fracture width, fracture network pattern, alter fracture propagation path, and proppant

placement. However, its role is still controversial because in many treatments the effect is

absent while in others is larger than expected (Vermylen and Zoback, 2011).

To sum up all the discussion, one can quote one sentence by Warpinski et al. (1993):

“The more complex the reservoir, the more complex the fracturing.”

2.5 SUMMARY

The following conclusions are drawn from above literature review and are now well

accepted in petroleum industry.

 The permeability of rock decreases with increasing confining pressure. The

reduction in permeability is 10 – 100 times larger for very tight rocks than for

permeable ones. One possible explanation of larger reduction in permeability is

smaller pore radii in tight formations. The flow capacity of an aperture depends

on a high power of its dimensions such fourth power for round capillaries and

third power for rectangular slits. Flow in tight gas sandstones (very low

31
permeability rocks) is mainly controlled by slit like apertures (micro cracks)

rather than round shape capillaries (Jones and Owens, 1980; Jennings et al., 1981;

Ostensen, 1983).

 The effect of confining pressure on pore volume compressibility and relative

permeability is not substantial and can be ignored for practical calculations.

 The reduction of permeability is further accelerated in presence of fractures, shale

streaks or other heterogeneity.

 The permeability reduction is always overestimated by conducting tests under

uniform hydrostatic loading in laboratory than non-uniform loading. True triaxial

tests at in situ conditions are better approximations.

 Pressure / stress dependent permeability must be incorporated in reservoir

engineering studies of tight formations because constant permeability assumption

for these formations is always misleading.

 Classical bi-wing planar fractures alone cannot explain fracture geometry and

productivity for unconventional reservoirs due to presence of natural and shear

fractures / fissures and geological discontinuities. Unconventional fracture models

are the more realistic and correct way of modeling complex fracture networks in

tight formations.

32
CHAPTER THREE: SUMMARY OF THE FIELD DATA

Wells under investigation are both dry gas wells of very low permeability completed in

the Cadomin formation. British Petroleum Canada, which was originally the participant

in the research consortium and agreed to provide the data, sold its Canadian gas holdings

to Apache Corporation who ultimately provided field data for this study. This chapter is

organized in several sections discussing basic data required for reservoir study,

stimulation treatment data from field, production data and geomechanical data. The

objectives of this chapter are to discuss and represent given data, highlight missing data,

and discuss methods used to generate necessary missing data by using best available

correlations from literature.

3.1 BASIC DATA

Basic data is representation of data that is required for reservoir study such as drainage

area, permeability, porosity, reservoir temperature, pressure, relative permeability, water

saturation and PVT data. Some of the above data was not available at the time of start of

the project so it was decided to use correlations to generate it. The following data was

missing from the information provided by the company:

 Drainage area

 Net pay thickness

 Connate water saturation

 PVT data

33
Drainage area of 10,000 x 6,000 ft2 (1377 acres) is assumed for current project. Initially

net pay thickness of 80 ft and connate water saturation of 25% was assumed based on

discussion with various sources. To generate PVT data for both wells the following step

by step process was followed using correlations.

 Correlation proposed by Sutton (1985) is used to calculate pseudo critical

properties of natural gas.

 Standing and Katz compressibility chart has been used for long time for natural

gas compressibility factor estimation. Several empirical correlations were

developed over years to calculate compressibility factor. Yarborough and Hall

(1974) presented an equation of state based on Carnahan-Starling equation of

state that accurately represents the Standing and Katz compressibility factor chart.

Dranchuck et al. (1974) developed a correlation for reduced density based on the

Benedict-Webb-Rubin type equation of state. Dranchuck and Abu-Kassem (1975)

derived an expression for reduced density that can be used to calculate

compressibility factor. All of the above three methods are applied to calculate z-

factor and it is observed that z-factor estimated by the above three methods give

essentially the same value for range of pressure of interest in this project. It was

then decided to use Yarborough and Hall (1973) method to calculate gas

compressibility factor.

 Carr et al. (1954) developed a graphical correlation for estimation of viscosity of

natural gas. Lee et al. (1966) proposed semi-empirical correlation in terms of gas

density, reservoir temperature, and molecular weight of gas for calculating

34
viscosity of natural gas. The viscosity of natural gas in this project is calculated

using the correlation presented by Lee et al. (1966).

Figure 3.1 represents natural gas compressibility factor and viscosity calculated by

correlations as discussed.

1.00 0.022
Compressibility factor (Z-factor)
Gas viscosity

0.97 0.019
Compressibility factor

Gas Viscosity (cp)


0.94 0.016

0.91 0.013

0.88 0.01
0 800 1600 2400 3200 4000
Pressure (psia)

Figure 3.1 Gas compressibility factor and viscosity of well A and B

The fluid and rock properties that are used as an input to reservoir simulator, TRS®

(provided by Taurus Reservoir Solutions Ltd.), are summarized in Table 3.1 for both

wells. As mentioned in the Introduction, field units system is used in this project unless it

is stated otherwise.

35
Table 3.1 Rock and fluid properties of well A and B (input to reservoir simulator)

Property Value (well A) Value (well B)


Total vertical depth, ft 5246 5308
Formation thickness, ft 80
Porosity, fraction 0.035 0.03
Permeability (Kh), mD 0.015 0.010
Permeability ratio (Kv/Kh), fraction 0.10
Reference pressure, psi 2000 2200
Reference depth, ft 5290 5350
Depth of gas oil contact, ft 15000
Depth of water oil contact, ft 15000
Formation compressibility, 1/psi 3.50E-06
Initial water saturation, percent 25
Reservoir temperature, °F 190
Gas specific gravity, fraction 0.613
Rate constraint, MMscf/d 7.0 5.5
Pressure constraint (minimum BHP), psi 300 275

3.2 STIMULATION DATA

Trican Well Service Ltd. performed the stimulation operations on these wells. One well

was fractured using hybrid fracturing technique, while the other was fractured by slick

water energized with CO2. Two different mesh sizes of proppant, i.e., 40/70 and 50/140

were used in each well in different quantities. Figures 3.2 and 3.3 represent the surface

and bottomhole proppant concentrations pumped in the injection field treatment. Figure

3.4 and 3.5 shows the bottomhole pumping rates and injection pressure in the two wells.

Treatment in Well B was pumped at higher rate and received larger total volume.
36
400
Inline - Well-A
Bottomhole - Well-A

300
Concentration (Kg/m3)

200

100

0
0 50 100 150 200 250
Elapsed Time (mins)

Figure 3.2 Proppant concentration of well A from field treatment report

300
Inline - Well-B
Bottomhole - Well-B
Concentration (kg/m3)

200

100

0
0 50 100 150 200 250 300
Elapsed Time (mins)

Figure 3.3 Proppant concentration of well B from field treatment report

37
15
Bottomhole Injection Rate - Well-A
Bottomhole Injection Rate - Well-B
12
Rate (m3/min)

0
0 50 100 150 200 250 300
Elapsed Time (mins)

Figure 3.4 Injection rates of well A and B from field treatment report

9000

Well A Well B
Bottomhole Injection Pressure (psi)

6000

3000
0 50 100 150 200 250 300
Time (mins)

Figure 3.5 Bottomhole field injection pressure - well A and B


38
The type, quantity and placement of proppant into formation play a very important role in

fracturing. Rushing and Sullivan (2003) concluded that slick water fractures performance

improves by using smaller proppant size (40/70 as opposed to 20/40) in Bossier tight gas

sands. Liu and Sharma (2005) proved through experimental study that proppant transport

is affected by the ratio of the proppant size to fracture width. The lateral flow of proppant

is significantly hindered as proppant size approaches the fracture width, particularly for

x-linked fluids, and proppant settling velocity is also reduced. This may eventually cause

tip screen-out or proppant bridging. The proppant is also distributed non-uniformly across

fracture width. Proppant particles are usually concentrated at the center of the flow

channel and thus travel faster than average fluid velocity because fluid velocity is

maximum at its center. Currently, it is believed in the industry that smaller proppant size

significantly improves initial production rates and effective fracture half length. The type

and amount of proppant used in each well is summarized in Table 3.2.

Table 3.2 Proppant types in well A and B

Property Value (well A) Value (well B)

40/70 mesh size sand, tonnes 50.6 30.0


50/140 mesh size sand, tonnes 31.3 63.2

3.3 PRODUCTION DATA

Average daily gas rate and cumulative gas production was the only complete data that

was available. Figure 3.6 and 3.7 represents the gas rate and cumulative production for

Well A and B respectively.

39
16000 1200
Avg daily Gas Rate
Cumulatve Gas Produciton

Cumulative Gas Production (MMscf)


12000 900
Avg Daily Gas Rate (Mscfd)

8000 600

4000 300

0 0
8-Jun-06 11-Dec-06 15-Jun-07 18-Dec-07 21-Jun-08 24-Dec-08 28-Jun-09 31-Dec-09

Figure 3.6 Field gas rate and cumulative production of well A

12000 1200
Avg daily Gas Rate
Cumulatve Gas Produciton

Cumulative Gas Production (MMscf)


9000 900
Avg Daily Gas Rate (Mscfd)

6000 600

3000 300

0 0
17-Aug-06 9-Feb-07 4-Aug-07 27-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 31-Dec-09

Figure 3.7 Field gas rate and cumulative production of well B


40
3.4 GEOMECHANICAL DATA

One of the most important requirements for geomechanics modeling is the knowledge of

in-situ stresses and of rock mechanics data, i.e., Poisson’s ratio, elastic modulus and

possibly other data (such as shear failure parameters). Data necessary to describe the

stress field is the variation of the principal stresses with depth and horizontal stress

orientation. It is well established that usually the vertical stress ( ) is larger than

maximum and minimum horizontal stresses ( and ) at large depth, although

exceptions may exist at shallow depths or in highly tectonic areas. The vertical stress at

depth (z) from the surface (in absence of tectonic events) is caused by the weight of the

overburden sediments and can be obtained as

where = Formation bulk density of the overburden, which can be obtained from well

logs

The propagation of hydraulic fractures and secondary fracture system is controlled by

minimum horizontal stress ( ) and the maximum stress. There are no reliable methods

for predicting in porous rocks without field data. The most common means of stress

orientation measurements are borehole breakouts, downhole telemetry, differential strain

analysis, anelastic strain relaxation, tiltmeter measurements, and microseismic

monitoring of hydraulic fractures. The methods used to measure stress magnitude are

micro-frac or mini-frac testing, and leak-off tests but estimates can also be obtained from

stress logs.

41
Anderson et al. (1973) proposed a relation to compute using Poisson’s ratio and

reservoir pore pressure.

where = Minimum horizontal stress

= Vertical stress

= Poisson’s ratio

= Biot’s constant

= Reservoir pore pressure

Biot’s constant plays an important role in poroelasticity.

where = = Solid grain (matrix) compressibility

= = Bulk compressibility

= Solid grain modulus

= Bulk modulus of grain or specimen

Newberry et al. (1985) modified Equation (3.2) for low porosity micofractured rocks.

Large amounts of data on fracture gradient have been compiled for US by Breckels and

Eekelen (1982) and for Canada by Bell and Babcock (1986). The stress orientation is

strongly affected by tectonic events so that is why data reported by authors show

considerable variations and deviations from the above equations. Therefore, log derived

stress data is more reliable for giving stress variation with depth rather than absolute

stress values of .

42
The is determined for this study from mini-frac analysis report and is estimated

from fracture study, reported by Apache Corporation, taking ratio of maximum to

minimum horizontal stress equal to 1.33 while vertical stress is determined from

calibrated density logs for current project. The horizontal stress values are also confirmed

from calibrated stress logs data. Static elastic modulus, Poisson’s ratio and uniaxial

compressive strength were measured in laboratory. Table 3.3 represents stress and rock

properties that are used in geomechanical modeling both for well A and B.

Table 3.3 Stress and rock properties for geomechanical modeling of well A and B

Property Value
Maximum horizontal stress gradient, psi/ft 1.13
Minimum horizontal stress gradient, psi/ft 0.85
Vertical Stress gradient, psi/ft 1.24
Formation density, kg/m3 2570
Static elastic modulus, psi 7.99E6
Poisson’s ratio 0.125
Uniaxial compressive strength, psi 46570
Friction angle, degrees 30
0.68
1.33

Bell et al. (1994) presented horizontal stress trajectories for western Canada sedimentary

basin in form of chart. The horizontal stress directions determined from breakouts for

western Canada sedimentary basin are given in Figure 3.8. It is assumed here that all

principal stresses are in direction of rectangular coordinate system. In our project well is

completed in x-direction which means that minimum horizontal stresses are also in this

direction. Maximum horizontal stresses are in y-direction while vertical stresses are in z-

direction in rectangular coordinate system.

43
Figure 3.8 Horizontal stress trajectories across western Canada sedimentary basin

(from Bell and Babcock, 1986 as reproduced by Mossop and Shetsen (1994))

44
CHAPTER FOUR: MODELING OF STATIC FRACTURES

The fractures created by hydraulic fracturing treatments, after being propped, do not

propagate with time, and are therefore known as static fractures. These fractures are

created to increase the productivity or injectivity of well. Static or stationary fractures are

filled with proppant injected at very high rate and pressure to create a conductive channel

in porous media of high porosity and high permeability. Permeability of this channel is

not constant and changes with time as stresses changes in wellbore / reservoir due to

pressure depletion. The properties of these highly conductive channels depend on the

type of stimulation treatments such as slickwater, x-linked gelled, hybrid, reverse hybrid,

acid, propped and un-propped fractures, proppant type, size, strength and quantity.

Palisch et al. (2010) defined slickwater treatment, which is also referred as waterfrac, as

“a fracture treatment that utilizes a large volume of water to create an adequate fracture

geometry and conductivity to obtain commercial production from low perm, large net pay

reservoirs.” The key benefits of slickwater frac job are reduced gel damage, lower costs,

and ability to reuse load and produced water, i.e., less environmental footprints

(Mayerhofer and Meehan, 1998; Sharma et al., 2004; Sharma et al., 2005; Liu et al.,

2006; Palisch et al., 2010) and complex fracture geometry (Palisch et al., 2010). Smaller

effective fracture half lengths due to proppant settling in low viscosity fluids, reduced

ability to transport proppant deeper into formation, large amount of water required for

proppant job, and narrower hydraulic fracture width are main concerns with slickwater

frac job (Rushing and Sullivan, 2003; Sharma et al., 2004; Sharma et al., 2005; Liu et al.,

45
2006; Palisch et al., 2010). Hybrid fracture treatment is not very different from slickwater

job except use of x-linked gels during the proppant pumping stage. The key benefits of

these types of job are better proppant transport which results in deeper placement of

proppant pack, larger fracture width (Sharma et al., 2005), and longer effective fracture

half length (Rushing and Sullivan, 2003; Sharma et al., 2004; Liu et al., 2006).

Susceptible gel damage, undesirable vertical fracture growth, and possibility of tip

screenout are major concerns tied to hybrid frac job (Sharma et al., 2005; Liu et al.,

2006). In reverse hybrid fracs, the fractures are created using a polymer gel pad followed

by slickwater for proppant placement (Sharma et al., 2005; Liu et al., 2006). Two

significant advantages of reverse hybrid frac job are larger fracture width allowing easier

proppant transport due to use of x-linked fluid pad and deeper placement of the proppant

into the formation due to viscous fingering as reported by (Sharma et al., 2005; Liu et al.,

2006). Fractures created by acid treatment are open channels and have very high porosity

(i.e., 1.0) while porosity of propped fractures is in the range of 0.32 – 0.40. The

permeability of an open fracture for single phase laminar flow between parallel plates is

Where = fracture width, ft

= Fracture permeability, mD

Turbulent flow, surface roughness, irregular surface shape, tortuosity, multiphase flow

and presence of different proppants all decrease fracture permeability and should be

considered in fracture permeability calculations. Settari et al. (1990) represented a

fracture in modeling productivity of fractured wells in single phase flow by introducing

46
transmissibility modifications of single grid system for both fracture and reservoir to

effectively minimize the grid effect and maximize the stability of numerical simulation.

The fundamental feature of the transmissibility modifications method is to combine the

fracture transmissibility with the transmissibility of reservoir blocks in the fracture plane,

without representing fracture by separate grid blocks. Ji et al. (2004a) discussed various

methods for modeling static fractures in reservoir simulation. According to their

conclusions transmissibility modification method can be used to model infinite and finite

conductivity fractures both in single and multiphase flow. The method can also be

extended to model dynamic fractures propagation in water injection, reservoir damage

and coupled geomechanical problems.

This chapter presents modeling of static fractures using concept of transmissibility

modifications in the reservoir simulator, TRS® provided by Taurus Reservoir Solutions

Ltd. Field injection data was analysed and used to model static fractures for production

modeling. The representation of fractures in reservoir simulator, grid refinement,

selection of fracture density, fracture spacing and estimation of fracture other parameters

will also be discussed in this chapter.

4.1 REPRESENTATION OF FRACTURES IN RESERVOIR SIMULATOR

4.1.1 Transmissibility Modifications

Transmissibility describes flow rate per unit pressure drop between two adjacent blocks.

Transmissibility has two parts, one part is flow properties part and other part is geometric

part. The flow part of transmissibility depends on the pressure and saturation of system.

47
As pressure changes in reservoir these properties changes and they are modified in

simulator accordingly. This part of transmissibility is also called the mobility of fluid.

From Darcy’s law

Where = Flow rate of phase l

= Effective fluid permeability = *

= Absolute permeability

= Relative permeability of phase l

A = Cross sectional area

= Viscosity of fluid of phase l

= Formation volume factor of phase l

= Change in pressure

= distance between centers of blocks between which the transmissibility is

computed

Mobility of fluid changes as pressure and saturation changes; this is automatically

updated in each time step by the simulator itself. Geometric part of transmissibility

remains constant if reservoir absolute permeability and grid size is not changing during

48
simulation. If permeability is considered stress/pressure dependent, this part of

transmissibility is also changing as stress/pressure changes. Here we will be discussing

only the geometric part of transmissibility for the case of a horizontal well with vertical

fractures. As the well is completed in x-direction, the one directional linear flow from

formation into the fracture will be in y-z plane. Transmissibility of non fractured blocks

is calculated by Equation (4.3). Transmissibility of blocks having fractures will be

modified using method given by Settari et al. (1990) and Ji et al. (2004a).

Substituting Equation (4.4) and (4.5) into (4.6)

Similarly for Z-direction,

Where = Transmissibility of fracture in y-direction

= Fracture area in y-direction =

= Fracture permeability

49
= Grid size in y-direction

= Matrix transmissibility in y-direction

= Matrix area in y-direction =

= Matrix permeability

= Fracture conductivity =

= Fracture width

= Transmissibility multiplier in y-direction

= Transmissibility multiplier in z-direction

Equation (4.8) and (4.9) are used to calculate transmissibility multipliers in y and z-

directions respectively and multipliers are represented in form of table in reservoir

simulator for fractured blocks.

4.1.2 Element of Symmetry Model

To reduce computational time, the line of symmetry for each well is taken by dividing

whole reservoir model into two halves in the x direction and performing the simulations

and analysis on a half model. Since fracture density is not constant for our project, the

line of symmetry is taken along the middle fracture by dividing the total number of

fractures into two halves. This implies that we must model only an odd number of

fractures (i.e., 3, 5, 7, etc.). The porosity, transmissibility and well index values for this

specific fracture are divided by 2 to take into account the symmetry. The output of the

simulator such as gas rates and cumulative production is then multiplied by factor of 2.0.

50
Another symmetry one could take an advantage of is in the y direction (solving only for

one wing of all the fractures), but this was not employed for our study.

4.1.3 Grid Refinement

Coarse grids are initially constructed for both wells, which are refined around each

fracture at later stage. Smaller grid sizes are chosen near the well bore to capture effects

of changing pressure and gas saturation, and larger grid sizes were used away from the

well. Grids are refined around each well completion (set of perforations) and fractures in

order to get better pressure profile near the vicinity of the well bore and of the fractures.

Irregular global refined grids of unequal size are used. Grid sizes in all directions i.e. x, y,

z - directions are changed from case to case to accommodate the fracture length due to

change in fracture density, net pay height and stimulated reservoir volume.

4.2 FRACTURE DENSITY

Fracture density plays an important part in production of multi fractured horizontal wells.

The larger is the number of active fractures, the higher will be the production. Initially we

were provided with very limited data about the frac jobs and assumed that a multi stage

fracturing job was done originally to create fractures. There was no information about

number of stages and details of the fracturing job. Therefore initially it was decided to

assume a total 9 numbers of fractures with assumed fracture spacing, which can be

changed at a later time. Fracture density has significant influence on production and

analysis of results and current fracture density may not be actual representation of in-situ

51
conditions, but the effects of fracture density on history matching will be discussed in

appropriate sections of subsequent chapters.

Later it was found that the wells were stimulated with a single treatment pumped into an

open hole. This was a very unusual, experimental type of treatment. Fracture spacing in

such a treatment is not controllable, but the microseismic monitoring showed that it was

successful in creating multiple fractures with spacing of about 50-60 m. This information

was used later to build the injection models in Chapter 7.

In the simulation models, the well is completed in x-direction in Cartesian co-ordinate

system. The orientation of the fractures from the geomechanical data given in Table 3.3 is

in y-z planes. The fracture height is growing in y-z plane; width in x-direction, while

length in x-y plane. Total length of horizontal section of wellbore is assumed 3300 ft

(1000m), which is an average wellbore length for horizontal wells in that area. The

schematic wellbore and fracture spacing is presented in Figure 4.1.

Figure 4.1 Well completion and fracture spacing – well A and B

52
Where Lf = Fracture half length

d = Average fracture spacing between 2 adjacent fractures, (assumed = 300 ft)

Line of symmetry is taken at fracture # 5 and reservoir model is constructed for a half of

the model comprising of ½ of fracture # 5 and fractures 6 to 9.

4.3 FRACTURES PARAMETERS ESTIMATION

Detailed modeling of fractures for each well based on wells injection data was beyond the

scope of this project. Modeling of fractures using correct amount and type of proppant,

fluid, type of treatment and injection rates and volumes is important and useful for

production analysis, but due to the constraints and uncertainty about the individual

fracture data in the open-hole treatment, it was decided to use simple equations for

fractures parameters estimations based on amount of proppant injected in each well. The

following assumptions are taken for fractures parameters calculations

 Fracture height is equal to pay zone or net pay thickness.

 Propped fracture width of 0.25 and 0.125 inches is assumed for well A and B

respectively.

 It was assumed that the proppants are contained within the main fractures only.

Complex fractures network of secondary fractures in stimulated reservoir volume

is assumed to be created, and these fractures will have conductivity (i.e., are

partially open) but will have no proppants in them.

 Porosity of each sand type is 37%.

 Density of both sand types is 2.65 gm/cc.

53
 Base Line permeability for each type of sand is taken from StimLab data sheet.

Values calculated based on closure stress and bottomhole temperature are given in

Table 4.1.

Table 4.1 Baseline permeability of sand used in well A and B

Property Mesh size Value (well A) Value (well B)

40/70 47800 42500


Permeability, mD
50/140 20000 18000

Sand of two different mesh types are used in each well. Permeability and conductivity of

each sand type in each well is calculated separately by assuming that amount of sand

injected downhole is contained only in the main fracture. The overall permeability and

conductivity of fractures is calculated based on fraction of sand injected in each particular

fracture in a well.

Putting Equation (4.13) into (4.12) and rearranging

Where = Volume of sand (downhole volume occupied by sand), ft3

54
= Conversion constant = 0.035315 (gm*ft3/kg*cc)

= Mass of sand in formation, Kg

= Density of sand, gm/cc

= Downhole volume, ft3

= Sand porosity

= Volume of all fractures, ft3

= Fracture width, inches

= Total fracture length in y-direction, ft

= Fracture half length, ft

= Number of fractures

= Fracture conductivity, mD-ft

= Baseline permeability of sand, mD

Sand of mesh size 40/70 in well A was used for the above sample calculation. The

calculated fracture parameters for each sand type both for well A and B are given in

Tables 4.2 and 4.3.

Sample Calculation for 40/70 type mesh sand in well A:

ft3

ft3

55
ft

= 35.68 ft

mD-ft

Table 4.2 Calculated fracture properties of sands in well A

Property – Well A Value – 40/70 Value – 50/140

Mass of sand in formation, Kg 50600 31300


Fraction of sand in formation 0.618 0.382
Baseline permeability of sand, mD 47800 20000
Downhole volume of sand, ft3 674 417
Total volume of all fractures, ft3 1070 662
Fracture half length, ft 35.68 22.07
Fracture conductivity, mD-ft 995.83 416.67
Closure stress, psi 3100
Fracture height, ft 80
Number of fractures (n) 9
Fracture width, inches 0.25
Sand porosity, fraction 0.37
Density of sand, gm/cm3 2.65
Overall average fracture permeability, mD 37175.58
Overall average fracture conductivity, mD-ft 774
Total fracture half length, ft 58

The overall fracture permeability is calculated by addition of permeability of different

sand types used in treatment. Permeability of the mixture is obtained by weighting

56
baseline permeability of each sand type by the fraction of that sand type in total injected

sand volume.

Table 4.3 Calculated fracture properties of sands in well B

Property – Well B Value – 40/70 Value – 50/140

Mass of sand in formation, Kg 30000 63200


Fraction of sand in formation 0.322 0.678
Baseline permeability of sand, mD 42500 18000
3
Downhole volume of sand, ft 340 842
3
Total volume of all fractures, ft 635 1337
Fracture half length, ft 42.31 89.13
Fracture conductivity, mD-ft 442.71 187.50
Closure stress, psi 3825
Fracture height, ft 80
Number of fractures (n) 9
Fracture width, inches 0.125
Sand porosity, fraction 0.37
3
Density of sand, gm/cm 2.65
Overall average fracture permeability, mD 25886.27
Overall average fracture conductivity, mD-ft 270
Total fracture half length, ft 131

4.4 SUMMARY

 Static fracture is modeled in reservoir simulator by combining the fracture

transmissibility with the transmissibility of reservoir blocks in the fracture plane,

57
without representing fracture by separate grid blocks. This method is known as

the transmissibility modification method.

 Refined grids are used near wellbore and around fracture blocks. Reservoir is

modeled by taking element of symmetry model.

 Total 9 fractures are assumed initially with average fracture spacing of 300 ft.

 Fracture parameters, i.e., half length, average permeability and conductivity are

calculated based on the amount of proppant injected into well and fracture.

58
CHAPTER FIVE: HISTORY MATCHING USING PSEUDO

CONTINUUM APPROACH – UNCOUPLED MODELS

History matching could be very simple or complex depending on nature of problem,

types and number of independent variables involves in it. There are no well defined

procedures and methods to get a history match. There is always more than one method by

which comparable results can be achieved. Then the question arises: which method to

choose and which to discard? The correct approach is to investigate the problem in detail

and put every possible scenario on the table for its solution. One should list all possible

and known variables that are involved and reduce their number by measured laboratory

and estimated data. Thorough investigation of remaining parameters and considering

extra physics of the problem, if needed, will also narrow down the number to a set of

variables that are critical to solution.

Different approaches to model the complex fracture geometry are being developed. Some

involve detailed solution of injection creating a single planar fracture (SPF), which then

interacts with the pre-determined network of fractures (Olson and Wu, 2012; Wu et al.,

2012). Another model is the creation of shear fracture network simultaneously with a

tensile fracture using pseudo-continuum technique (Nassir et al., 2012) or displacement

discontinuity methods (Tao et al., 2009). The approach considered in this study assumes

that the creation/enhancement of fracture system can be represented by pseudo-

continuum with dynamically changing permeability (as a tensor) as a function of

effective stress. The matrix permeability can also be made a function of effective stress

59
and all these functions have a hysteresis in the injection-production cycle. This method is

an extension of the techniques used in tight gas (Settari et al., 2002b, 2009) and has been

applied to Eagleford, Barnett and other shale plays. This method can be used for the

optimization of the treatment because it includes modeling the creation of the stimulated

reservoir volume as well as production modeling. It can be used to optimize the number

of stages and sets of perforations within the stage, treatment size for a stage, use of

different fluids, etc. It has been also used to investigate effects of stress shadowing, well

spacing, and can be further simplified to an uncoupled modeling. In this Chapter of the

thesis, we do not deal with the simulation of the fracturing (stimulation process) itself.

Therefore the size of the SRV and the induced fracture inside each SRV are matching

parameters.

This chapter presents investigation of the role of fracture conductivity, net pay thickness,

permeability multipliers and stimulated reservoir volume on gas rate and cumulative

production. Simulations are run for different cases and results are then compared with

field data. The sensitivity of the results to any particular parameter and deviation of

results from field data are discussed in each section. Assumptions made for some cases to

simplify problem are discussed in that particular section. Figure 5.1 represents flow chart

of simulation runs performed in this chapter. Every previous case serves as base case for

next simulation run.

60
Field Production History Matching

Baseline Propped Fracture Permeability


Case – 5.1.1
Infinite Fracture Conductivity
– An Analytical Approach
Realistic Propped Fracture Permeability
Case – 5.1.2

Effect of Net Pay Thickness (NPT)


Case – 5.2

Effects of Stimulated Reservoir Volume (SRV)

History Matching using Manual Change of


Reservoir Permeability with Time

Figure 5.1 Flow chart of simulation cases – Chapter 5

5.1 ROLE OF FRACTURE CONDUCTIVITY

The conductivity of fracture can be calculated from fracture permeability and width, and

it is represented by the following equation.

Where = Fracture conductivity

= Fracture permeability

= Fracture width

Since in this project fractures are propped, fracture conductivity is controlled by proppant

permeability and propped fracture width. Propped fracture permeability and width are

61
referred to as fracture permeability and fracture width respectively throughout this thesis

unless mentioned otherwise. American Petroleum Institute (API) developed standard

procedures for measuring the conductivity of proppants and these procedures are

documented in API RP-61. Later Much and Penny (1987) recommended minor changes

in API test that resulted in significant improvements in results, which was approved by

International organization for Standardization (ISO) in 2006 as ISO 13503-5. Palisch et

al. (2007) presented comparison between the two tests for high and low quality ceramic

proppant and reported that loss in conductivity measured by modified test was as much as

85% depending on proppant quality and test conditions. Industry has been using the API

modified test since that time and the test is usually referred as “long term” conductivity

test. “Short term” conductivity test is referred to the original procedure of RP-61.

Baseline or Laminar or Reference conductivities reported by service companies and

proppant manufacturers are measured using the long term tests.

5.1.1 Baseline Propped Fracture Conductivity

For our wells, baseline conductivity is calculated using Equation (5.1) by taking proppant

permeability from StimLab database, and assuming fracture width. Case 5.1.1 in Tables

5.1 and 5.2 represents the case where baseline conductivity for fracture is used in

simulation for wells A and B. The other properties in Tables 5.1 and 5.2 are calculated

using equations given in Chapter 4.

62
Table 5.1 Fracture properties of Case 5.1.1 – 2 for well A

Values Values
Property – Well A
(Case 5.1.1) (Case 5.1.2)
Net pay thickness, ft 80
Fracture width, inches 0.25
Fracture half length, ft 58
Fracture height, ft 80
Fracture conductivity, mD-ft 774 39
25081 1301
250081 13001

Table 5.2 Fracture properties of Case 5.1.1 – 2 for well B

Values Values
Property – Well B
(Case 5.1.1) (Case 5.1.2)
Net pay thickness, ft 80
Fracture width, inches 0.125
Fracture half length, ft 131
Fracture height, ft 80
Fracture conductivity, mD-ft 270 13
13501 651
135001 6501

5.1.2 Realistic Propped Fracture Conductivity

Even though the modified API test (using increasing testing time and variable

temperatures) gives good indication of proppant performance, it still largely

overestimates conductivity values. Several authors have worked in this area and

investigated the factors that could potentially decreases conductivity, which is critical for

reservoir simulation of productivity. Barree et al. (2003), Dedurin et al. (2006), and

Palisch et al. (2007) estimated realistic fracture conductivity by investigating the factors

63
that led to conductivity reduction and explained it in detail. The following are some

factors that according to these authors play a major role in defining actual fracture

conductivity and should be taken in to account in fracturing design treatment.

 Non Darcy flow

 Multiphase flow

 Relative permeability effects

 Proppant concentration

 Gravity segregation and capillary effects

 Filter cake at fracture faces

 Gel damage through proppant pack

 Proppant embedment into fracture face

 Spalling of formation into fracture

 Fines migration

 Cyclic stresses

 Sieve size distribution and proppant types

 Long term proppant degradation, non-uniform proppant distribution and elevated

temperature effects

It has been shown that all these effects (which are not considered in API test procedure)

are cumulative and can reduce fracture conductivity by as much as 99% under realistic

conditions. Schubarth et al. (2006) supported these conclusions in their study.

Montgomery and Steanson (1985) stated that four factors, namely propped fracture area,

propped fracture conductivity, reservoir permeability and drainage area mainly control

64
improvements in productivity. They showed that conductivity degradation continues with

extended time up to 9 months. Conductivity of proppants was continuously reducing until

it reached equilibrium due to particle crushing and rearrangements. Different proppants

may be affected by the above factors in a different way but all of them typically

experience 2 – 3 orders of magnitude of reduction in conductivity. Therefore, estimating

realistic fracture conductivity is the key to realistic predictions of stimulation

effectiveness.

In light of the above discussion, it was decided to use multiplier of 0.05 to take all this

factor into account, i.e., fracture conductivity was reduced by 95%. Case 5.1.2 in Tables

5.1 and 5.2 represent the case where realistic fracture conductivity is used in simulation

for wells A and B by using 5% of baseline fracture conductivity. Transmissibilities for

fracture flow (see chapter 4) are calculated for this case using the modified values of

conductivity. Simulations were run for both cases, i.e., Case 5.1.1 and 5.1.2 and results

are presented in Figures 5.2 and 5.3 for both wells.

65
16 1200
Field Avg Gas Rate Gas Rate - C-5.1.1
Gas Rate - C-5.1.2 Cumulative Gas - C-5.1.1
Cumulative Gas - C-5.1.2 Field Cumulative Gas

Cumulative Gas Prodcution (MMscf)


12 900
Avg Daily Gas Rate (MMscfd)

8 600

4 300

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 5.2 Comparison of field data and simulation results of Cases 5.1.1 - 2 - well A

12 1200
Field Avg Gas Rate Gas Rate - C-5.1.1
Gas Rate - C-5.1.2 Field Cumulative Gas
Cumulative Gas - C-5.1.1 Cumulative Gas - C-5.1.2

Cumulative Gas Production (MMscf)


9 900
Avg Daily Gas Rate (MMscfd)

6 600

3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 5.3 Comparison of field data and simulation results of Cases 5.1.1 - 2 - well B
66
The predicted gas rate and cumulative gas is much smaller than field results for both

wells, but the difference between the results of the two cases is very small for well A

while it is a bit larger for well B. Since fracture width for both wells is kept constant

throughout and only permeability is changed, it means that changing proppant

permeability does not affect the results significantly. It can be concluded from these runs

that fracture for current project are acting almost as “infinite conductivity fractures” (the

term will be discussed next). Sharp decline in gas rate in early time for either well is due

to large pressure drop near the wellbore and around fractures.

5.1.3 Infinite Fracture Conductivity – An Analytical Approach

Dimensionless fracture conductivity or capacity, Fcd, is the ratio of the ability of the

fracture to transmit fluids down to fracture and into wellbore to the ability of the

formation to deliver fluids into the fracture. It is defined as

Where = Dimensionless fracture conductivity

= Fracture permeability

= Fracture width

= Formation permeability

= Fracture half length

Dimensionless fracture conductivity has been in use since the earliest days of hydraulic

fracturing. McGuire and Sikora (1960) plotted productivity index against relative

67
conductivity, dimensionless fracture conductivity, for various fracture half length. They

showed that productivity increases 2 to 6 times in range of 103 – 104 of relative

conductivity while the increase is negligible for larger values of relative conductivity.

Prats (1961) concluded that the effective wellbore radius is always equal to a half of

fracture half length for an infinite conductivity fracture and for Fcd > 10 increase in

effective wellbore radius is small and fracture behaves as an infinite conductive fracture.

Dedurin et al. (2006) in their paper confirmed his conclusions. It is typical that

dimensionless fracture conductivity has higher values than predicted by Prats for steady

state production. Cinco-Ley and Smaniego-V (1981) performed pressure transient

analysis on vertically fractured well producing at constant flow rate in an infinite,

isotropic, and homogenous reservoir and showed that for dimensionless time greater than

1, the Fcd value must be 30. They and Barree et al. (2003) argued that for Fcd > 30

fractures can be represented as an infinite conductivity fractures. Pearson (2001)

suggested Fcd value of 10 for infinite conductivity based on Prats’ work and industrial

experience and concluded that realistic conductivity should be utilized when calculating

Fcd. It is an important parameter in the design of hydraulic fracture treatments and it is a

tool that helps engineers in predicting benefits of fracture simulation.

An infinite conductivity fracture of half-length Lf can be, in pseudo-steady state

conditions, replaced by an effective well bore radius or skin using the following relation,

Skin from Equation (5.3)

68
Where = Effective wellbore radius

= Wellbore radius

= Skin

Positive skin effect means there is an increase in pressure drop, which indicates extra

flow resistance near the wellbore while negative skin effect means decrease in pressure

drop, which is characterized by flow enhancement near wellbore. Effective wellbore

radius increases with negative skin, i.e., hydraulic & acid fracturing while it decreases

with positive skin, i.e., mud invasion, drilling fluid loss, and near wellbore formation

damage.

Dimensionless fractures conductivity concept will be used here to plot dimensionless

effective wellbore radius against Fcd. It is proved from simulation results that fractures

are almost infinite conductive fractures. Criterion suggested by Prats, Dedurin and

Pearson (i.e., fractures are of infinite conductivity if Fcd ≥ 10) is used here to analytically

predict whether fractures have finite or infinite conductivity. For this purpose

dimensionless fracture conductivity, effective wellbore radius, and skin is calculated

using Equations (5.2, 5.3 and 5.4) and is given in Table 5.3 for baseline and realistic

fracture conductivity cases for both wells. The dimensionless effective wellbore radius is

plotted against dimensionless fracture conductivity in Figure 5.4.

Table 5.3 Dimensionless fracture conductivity of Cases 5.1.1 - 2 for well A and B

Property Values – Well A Values – Well B

Wellbore radius, ft 0.25


Fracture width, inches 0.25 0.125
Reservoir permeability, mD 0.015 0.010

69
Fracture half length, ft 58 131
Case 5.1.1
Fracture permeability, mD 37175.60 25886.30
Skin, - 4.75 - 5.56
Dimensionless fracture conductivity 890 205
Case 5.1.2
Fracture permeability, mD 1858.80 1294.32
Skin, - 4.75 - 5.56
Dimensionless fracture conductivity 44.51 10.30

Well A Well B
1
Dimensionless effective wellbore radius
(rwe/Lf)

0.1

0.01
0.01 0.1 1 10 100 1000
Dimensionless fracture conductivity (Fcd)

Figure 5.4 Dimensionless fracture conductivity of Cases 5.1.1 - 2 – Well A and B

Dimensionless effective wellbore radius for each well is equal to 0.5 and dimensionless

fracture conductivity is large than 10 for all cases. This analysis confirms our earlier

conclusion that fractures in this project is infinite conductivity fractures. Case 5.1.2 for

both wells will be the base case for next simulation runs.

70
5.2 EFFECT OF NET PAY THICKNESS (NPT)

Net pay thickness used in earlier cases was an assumed value of 80 ft discussed before.

Sponsor company provided us with more information on well and geology in a recent

meeting. It was decided to run a case using updated value of net pay thickness of 100 ft.

Increasing net pay thickness means larger pore volume and eventually higher production.

Case 5.2 shown below is considered as base case for further cases and investigation.

Summary of fracture parameters and transmissibility multipliers both for well A and B

for Case 5.2 is given in Table 5.4.

Table 5.4 Fracture properties of Case 5.2 for well A and B

Property (Case 5.2) Values – Well A Values – Well B


Net pay thickness, ft 100
Fracture height, ft 100
Fracture width, inches 0.25 0.125
Fracture half length, ft 46 105
Fracture conductivity, mD-ft 39 13
1301 651
13001 6501

71
16 1200
Field Avg Gas Rate Gas Rate - C-5.1.2
Gas Rate - C-5.2 Cumulative Gas - C-5.1.2
Cumulative Gas - C-5.2 Field Cumulative Gas

12 900

Cumulative Gas Prodcution (MMscf)


Avg Daily Gas Rate (MMscfd)

8 600

4 300

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 5.5 Comparison of field data and simulation results-Cases 5.1.2 & 5.2 -well A

12 1200
Field Avg Gas Rate Gas Rate - C-5.1.2
Gas Rate - C-5.2 Field Cumulative Gas
Cumulative Gas - C-5.1.2 Cumulative Gas - C-5.2

Cumulative Gas Production (MMscf)


9 900
Avg Daily Gas Rate (MMscfd)

6 600

3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 5.6 Comparison of field data and simulation results-Cases 5.1.2 & 5.2 - well B

72
Figures 5.5 and 5.6 represent comparison of field data and simulation results of current

case for well A and B respectively. Production in early time is mostly from the fracture

and therefore base case and current case results are similar in early time. Increase in net

pay thickness means larger pore volume hence more gas in place and larger production

compared to the base case.

5.3 EFFECTS OF STIMULATED RESERVOIR VOLUME (SRV)

Hydraulic fracturing technology has improved immensely since the adoption of this

process. The full understanding and optimization of hydraulic fracture simulation will not

be possible until accurate methods are available to directly image the fracture

characteristics. The only methods available for in-situ imaging are the use of downhole

receiver and tiltmeter arrays for detection and monitoring of fracture induced

microseismic events and deformations. Microseismic events (MS) are primarily shear

movements of a limited area of rock surface. Whether MS can also detect tensile events is

still somewhat controversial. A pure shear movement emits both compressional waves (p-

waves) and shear waves (s waves) and detection of p-wave and s-wave arrivals at

observation well provide the information to accurately locate the source point of the

event. A 3-dimensional location for any particular event can be found using single

vertical array of receivers and locations of all these events are then used to generate an

image of a growing fracture. Exact understanding of causes of these events which is

necessary to interpret the measurements is still evolving as these shear failure events can

be found hundreds of feet away from fractures and even outside the stimulated zone.

73
It is now believed that apart from major fractures there is also a network of small

fractures and secondary fractures being created during hydraulic fracturing treatments.

The leak-off of fracturing fluid into the formation perpendicular to fracture face is

contributing to their development, but they can extend beyond the fluid penetration. The

region in which the secondary fracturing was created is often called Stimulated Reservoir

Volume (SRV). These secondary fractures could be the opening of already existed

natural fractures in the formation or it could be newly created network of open fractures

due to shear failure of rock. These secondary fractures are highly conductive open

channels because these are open fractures. It is possible that early time production of our

wells could be due to the contribution of main fracture and network of these open

fractures and late time production curves are comparatively flat due to the closure of

these fractures.

To measure the effects of these secondary fractures on production, few cases assuming

the presence of SRV were run and results were then compared with base case. The

volume around the main fracture and wellbore where open secondary fractures are

concentrated (i.e., the SRV) is shown schematically on Figure 5.7. In left side (a) of the

figure the schematic of fracture and SRV created downhole in reservoir is shown, while

the right side (b) shows the representation of the fracture and SRV by a rectangular

region in reservoir simulator. To model increase of permeability inside SRV, a multiplier

which we will call an initial permeability multiplier (IPM) is used. It means that

permeability of rock inside SRV is enhanced by some factor due to fracture development.

This permeability can then be decreased as a function of the depletion of the reservoir, as

will be discussed in Chapter 6.

74
Figure 5.7 Stimulated reservoir volume (SRV) schematic – Complex fracture

geometry representation

Calculated transmissibility and assumed values of SRV and initial permeability

multipliers for all sub-cases in each well are presented in Table 5.5. 6,18 25, 35, and 40 in

Table 5.5 represent IPM and numbers in bracket followed IPM values represent modified

reservoir permeability after multiplying IPM. The method of transmissibility calculation

and formulation has been given in chapter 4.

Table 5.5 SRVand initial permeability multiplier of Cases 5.3.1 – 6 for well A and B

Property (Cases 5.3.1 - 6) Values – Well A Values – Well B


Case 5.3.1
218 109
2168 1084
SRV (one wing), ft3 40x3000x100 40x3000x100
Initial permeability multiplier 6 (0.09, 0.009) 6 (0.06, 0.006)
Case 5.3.2
218 109
2168 1084

75
SRV (one wing), ft3 80x3000x100 80x3000x100
Initial permeability multiplier 6 (0.09, 0.009) 6 (0.06, 0.006)
Case 5.3.3
218 109
2168 1084
SRV (one wing), ft3 100x3000x100 100x3000x100
Initial permeability multiplier 6 (0.09, 0.009) 6 (0.06, 0.006)
Case 5.3.4
73 20
723 187
SRV (one wing), ft3 120x3000x100 140x3000x100
Initial permeability multiplier 18 (0.27, 0.027) 35 (0.35, 0.035)
Case 5.3.5
73 20
723 187
SRV (one wing), ft3 120x300x100 140x300x100
Initial permeability multiplier 18 (0.27, 0.027) 35 (0.35, 0.035)
Case 5.3.6
53 17
521 164
SRV (one wing), ft3 120x300x100 140x500x100
Initial permeability multiplier 25 (0.375, 0.0375) 40 (0.06, 0.006)

Comparison between simulation results and field data for Cases 5.3.1 to 5.3.6 are

presented in Figure 5.8 and 5.9 for well A and in Figure 5.10 and 5.11 for well B.

76
16 1600
Field Avg Gas Rate Gas Rate - C-5.2
Gas Rate - C-5.3.1 Gas Rate - C-5.3.2
Gas Rate - C-5.3.3 Field Cumulative Gas
Cumulative Gas - C-5.2 Cumulative Gas - C-5.3.1
Cumulative Gas - C-5.3.2 Cumulative Gas - C-5.3.3

Cumulative Gas Prodcution (MMscf)


12 1200
Avg Daily Gas Rate (MMscfd)

8 800

4 400

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 5.8 Comparison of field data and simulation results of Cases 5.3.1 - 3 - well A

16 2800
Field Avg Gas Rate Gas Rate - C-5.3.4
Gas Rate - C-5.3.5 Gas Rate - C-5.3.6
Field Cumulative Gas Cumulative Gas - C-5.3.4
Cumulative Gas - C-5.3.5 Cumulative Gas - C-5.3.6
12 2100

Cumulative Gas Prodcution (MMscf)


Avg Daily Gas Rate (MMscfd)

8 1400

4 700

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 5.9 Comparison of field data and simulation results of Cases 5.3.4 - 6 - well A
77
12 1500
Field Avg Gas Rate Gas Rate - C-5.2
Gas Rate - C-5.3.1 Gas Rate - C5.3.2
Gas Rate - C-5.3.3 Field Cumulative Gas
Cumulative Gas - C-5.2 Cumulative Gas - C-5.3.1
1200
Cumulative Gas - C-5.3.2 Cumulative Gas - C-5.3.3
9

Cumulative Gas Production (MMscf)


Avg Daily Gas Rate (MMscfd)

900

600

3
300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 5.10 Comparison of field data and simulation results of Cases 5.3.1 - 3-well B

12 2800
Field Avg Gas Rate Gas Rate - C-5.3.4
Gas Rate - C-5.3.5 Gas Rate - C5.3.6
Field Cumulative Gas Cumulative Gas - C-5.3.4
Cumulative Gas - C-5.3.5 Cumulative Gas - C-5.3.6

Cumulative Gas Production (MMscf)


9 2100
Avg Daily Gas Rate (MMscfd)

6 1400

3 700

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 5.11 Comparison of field data and simulation results of Cases 5.3.4 - 6 -well B

78
We will first discuss the first three cases, i.e., Cases 5.3.1 - 3 given in Figure 5.8 for well

A and in Figure 5.10 for well B. In these three cases the width of the SRV is varied from

40 to 100 ft while the same initial permeability multiplier is used to enhance permeability

of reservoir rock inside SRV. The results of these cases are quite similar in early time but

difference is quite large for late time especially for the first two cases. The flow regions

can be divided into three segments. Region I represent is flow through propped fracture

that transfers fluid into wellbore, region II is the SRV which is transmitting reservoir

fluids to fractures, and region III is the flow in the reservoir itself that is feeding to the

SRV. The 2-dimensional top view of all the three regions around a fracture is shown in

Figure 5.12.

Figure 5.12 Areal view of fracture in x-y plane – Flow pattern representation

By definition of hydraulic diffusivity, the greater is the permeability, the larger is the rate

of pressure propagation. Pressure propagation will be faster in region I and II and

comparatively slower in region III. Early time higher production is mainly from SRV.

79
When pressure propagation reaches SRV boundary, then production declines due to

smaller reservoir permeability. The time at which production curves deviate from each

other, such as between Cases 5.3.2 and 5.3.3, indicates the time taken by pressure to

reach the boundary of region III. These cases give us some indication that for higher early

time production, we either need larger initial permeability multiplier or larger SRV width

or combination of both.

Case 5.3.4 is run to achieve early time history match by varying both width of SRV and

initial permeability multiplier. Early time history match is achieved with parameters

given in Table 5.5 and results are in Figure 5.9 and 5.11 both for well A and B

respectively. The increase in permeability multiplier for well A is 3 times and 6 times for

well B as compared to Case 5.3.3 and with little increase in SRV width. Now the task is

to select an optimum length of SRV and also match the results at late time. For this

purpose two more cases are run, i.e., Case 5.3.5 and 5.3.6 by varying both length of SRV

and initial permeability multiplier. The summary of these cases is given in Table 5.5

while results for those cases are presented in Figure 5.9 for well A and 5.11 for well B.

All the parameters in Case 5.3.4 and 5.3.5 are same except for the length of SRV but

there is huge difference in results. The same argument as discussed before can be applied

here to explain the difference in results and decrease in production. If one analyzes the

final case, i.e., Case 5.3.6 for any well, it can be said that although results are improved in

early time, simulation over predicts results at late time. The reason is that the

permeability multiplier is constant throughout the simulation, but in reality permeability

inside the SRV and outside in the formation decreases with reservoir pressure depletion

80
(or with increase in net effective stress). Thus permeability at late time is much smaller

than permeability in early time due to reservoir deformation.

It is therefore concluded that history match with field data is not possible by using

constant permeability multiplier or constant permeability overall. Pressure / stress

dependent permeability is necessary to obtain a history match, which will be the topic of

discussion in subsequent chapters. As Case 5.3.6 provide a reasonably good match with

field data, this case will be used as a base case for further runs.

5.4 HISTORY MATCHING USING MANUAL CHANAGE OF RESERVOIR

PERMEABILITY WITH TIME

This case is run to authenticate our conclusions of need for using pressure / stress

dependent permeability. There are different ways to change permeability with respect to

time. One of them is to use time dependent transmissibility. As permeability is contained

in the transmissibility equation, modifying transmissibility while keeping other

parameters constant means modifying permeability with respect to time. Field data

analysis also gives us some indications that permeability is initially higher and decreases

continuously with time. It is vital to first test this concept, which can be done in the TRS

software by changing the permeability at specified time intervals manually. Different

permeability multiplier, value smaller than 1, is applied both inside and outside SRV at

specified time intervals to get history match, which is represented in Figure 5.13 for Well

A. The true pressure / stress dependent permeability effects will be considered and

modeled in Chapter 6.

81
Table 5.6 Fracture properties of Case 5.4 for well A and B

Property (Case 5.4) Values – Well A Values – Well B


Net pay thickness, ft 100
Fracture height, ft 100
Total number of fractures 9
Fracture width, inches 0.25 0.125
Fracture half length, ft 46 105
Fracture conductivity, mD-ft 39 13
SRV (one wing), ft3 120x300x100 140x500x100
Initial Permeability Multiplier 25 (0.375, 0.0375) 40 (0.06, 0.006)

Figure 5.14 and 5.15 represent comparison of field data and results from simulation for

current case both for well A and B respectively. It is apparent that it is possible to

duplicate the history very accurately. This method is of course purely empirical, but

provides motivation for including geomechanics.

1
Inside SRV

outside SRV
0.8
Permeability Multipliers

0.6

0.4

0.2

0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 5.13 Time dependent permeability multipliers-History matched case–Well A

82
16 1200
Field Avg Gas Rate Gas Rate - C-5.4

Cumulative Gas - C-5.4 Field Cumulative Gas

Cumulative Gas Prodcution (MMscf)


12 900
Avg Daily Gas Rate (MMscfd)

8 600

4 300

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 5.14 History matching of well A using time dependent permeability

12 1200
Field Avg Gas Rate Gas Rate - C-5.4

Field Cumulative Gas Cumulative Gas - C-5.4

9 900

Cumulative Gas Production (MMscf)


Avg Daily Gas Rate (MMscfd)

6 600

3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 5.15 History matching of well B using time dependent permeability


83
5.5 SUMMARY

 Using baseline conductivity in simulation and other calculations will always

overestimate results; therefore realistic proppant conductivity should be used.

 Fractures have infinite conductivity for both wells for this project, as it was

proved both by running simulations and comparing production curves and

dimensionless fracture conductivity calculations.

 Net pay thickness has no effects on production in early time but production

increases at late time due to larger reservoir size.

 Stimulated reservoir volume has significant effects on results. Permeability inside

SRV must be larger than formation permeability by factor greater than 2.0.

 Early time history match can be achieved by using permeability multiplier in SRV

and constant permeability in outside formation but it is difficult to match the late

time results using the same parameters.

 History match is impossible without the use of pressure / stress dependent

permeability. The last case run verified this conclusion and it proved (from results

in Figure 5.14 and 5.15) that history match is only possible by introducing some

form of time changing permeability.

It has been established in this chapter that history match without the use of pressure /

stress dependent permeability is not possible. History matching using pressure / stress

dependent permeability using both uncoupled reservoir and a fully coupled

geomechanical model will be discussed in detail in Chapter 6.

84
CHAPTER SIX: HISTORY MATCHING AND MODELING OF

GEOMECHANICAL EFFECTS USING PRESSURE / STRESS

DEPENDENT PERMEABILITY

It was proved and concluded in Chapter 5 that history match is not possible without the

use of time dependent transmissibility or permeability. A function is needed that defines

permeability as function of effective stress or pressure. Stress dependent permeability

always means effective stress dependent permeability in this thesis and pressure

dependent permeability means that the independent variable is pressure, which is

equivalent to the assumption that total stresses are constant. Average (mean) effective

stress is defined as

where = Mean effective stress

= Total mean stress

= Biot’s constant

= Fluid pressure

= Vertical stress

= Maximum horizontal stress

= Minimum horizontal stress

85
In Equation (6.1) is part of total mean external stress carries by solid matrix

(framework) and the remaining part, , is carried by the fluid. The equation can also

be written for one dimensional system in which mean stress will be replaced by

corresponding stress values. Permeability is in general a function of effective stress

which means that it is function of both total stress and fluid pressure. However, for

simulation it can be also treated approximately as a function of total stress only (which is

rare) or fluid pressure only depending on which is kept as an independent variable.

where = Permeability

= Initial total mean stress

Equation (6.3) represents a case where permeability is function of both total mean stress

and fluid pressure and in (6.4) it is a function of fluid pressure only.

This chapter presents the correlation used to generate pressure / stress dependent

permeability data, history matching using this data, stress sensitivity of formation, effect

of shape of pressure / stress dependent permeability multipliers on results, comparison of

results between uncoupled and fully coupled geomechanical models and approximation

of geomechanical effects in conventional reservoir simulator using uniaxial deformation

theory. Cases run for detailed investigation of various parameters and results are

compared to field data. The conclusions and discussions are presented at end of each

particular section. Figure 6.1 represents flow chart of simulation runs performed in this

chapter. Every previous case serves as base case for next simulation run.

86
Field Production History Matching using
Pressure / Stress Dependent Permeability

Case – 6.1 – 15 Fractures – Uncoupled


Production Models

Effects of Shape of Pressure Dependent


Permeability Curves – Uncoupled
Production Models – Cases – 6.3.1.1–4

Effects of (SRV) – Uncoupled


Production Models – Cases – 6.3.2
ase – 5.2

History Matching using Pressure History Matching using Pressure


Dependent Permeability outside SRV Dependent Permeability both Inside
Cases – 6.3.3.1 – 2 and Outside SRV – Cases – 6.3.5

History Matching using Stress Dependent Permeability both Inside and


Outside SRV – Coupled Production Geomechanical Models – Constant
Mean Total Stress Cases (Case – 6.5)

Approximation of Geomechanical Effects in Conventional


Reservoir Simulator using Uniaxial Deformation Theory

History Matching using Stress Dependent Permeability both Inside and


Outside SRV – Uncoupled and Coupled Production Geomechanical
Models – Variable Mean Total Stress Cases (Case – 6.6)

Figure 6.1 Flow chart of simulation cases – Chapter 6

87
6.1 DISCUSSION ON NEW INFORMATION

In a recent meeting with sponsored company representatives some more information

about in-situ fracture density and fracture spacing was given. According to microseismic

report from Apache Corporation, average in-situ fracture spacing is 55 m (180 ft) while

fracture half-length is 250 m (820 ft). It was decided to rework all the fracture parameters

calculation shown in Chapter 4 in light of the new data. For the same amount of volume,

the larger the number of fractures, the smaller is fracture length given by Equation (4.14).

A new base case to compare other runs with in this Chapter was created by choosing

Case 5.3.4 as a starting point (the reason for choosing this case is early time relatively

good match). Changing fracture density changes fracture half length and as a result

smaller initial permeability multiplier is needed to get the same results as Case 5.3.4. The

purpose of this new base case is to get the same results as the Case 5.3.4, so that the

current base case uses new updated information. Calculated rock and fracture properties

that are used as input in reservoir simulator for the new base case 6.1 are summarized in

Table 6.1 and 6.2 for well A and B respectively.

Table 6.1 Fracture properties of Case 5.3.4 and 6.1 for well A

Values Values
Property – Well A
(Case 5.3.4) (Case 6.1)
Net pay thickness, ft 100
Fracture height, ft 100
Fracture width, inches 0.25
Fracture conductivity, mD-ft 39
SRV (one wing), ft3 120x3000x100
Fracture half length, ft 46 28

88
Initial permeability multiplier 18 (0.27, 0.027) 10 (0.15, 0.015)
Number of fractures, n 9 15
73 131
723 1301

Table 6.2 Fracture properties of Case 5.3.4 and 6.1 for well B

Values Values
Property – Well B
(Case 5.3.4) (Case 6.1)
Net pay thickness, ft 100
Fracture height, ft 100
Fracture width, inches 0.125
Fracture conductivity, mD-ft 13
3
SRV (one wing), ft 140x3000x100
Fracture half length, ft 105 63
Initial permeability multiplier 35 (0.35, 0.035) 12 (0.12, 0.012)
Number of fractures, n 9 15
20 55
187 543

Figures 6.2 and 6.3 represent the comparison of the old base case and Case 6.1 and also

with field data.

89
16 2800
Field Avg Gas Rate Gas Rate - C-5.3.4 Gas Rate - C-6.1
Field Cumulative Gas Cumulative Gas - C-5.3.4 Cumulative Gas - C-6.1

Cumulative Gas Prodcution (MMscf)


12 2100
Avg Daily Gas Rate (MMscfd)

8 1400

4 700

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.2 Comparison of field data and simulation results of Case 6.1 - well A

12 2800
Field Avg Gas Rate Gas Rate - C-5.3.4 Gas Rate - C-6.1
Field Cumulative Gas Cumulative Gas - C-5.3.4 Cumulative Gas - C-6.1

Cumulative Gas Production (MMscf)


9 2100
Avg Daily Gas Rate (MMscfd)

6 1400

3 700

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.3 Comparison of field data and simulation results of Case 6.1 - well B

90
6.2 PRESSURE / STRESS DEPENDENT PERMEABILITY CORRELATION

The permeability dependence of rock and fractures on stress and fluid pressure in pores is

well documented in literature. The permeability of oil or gas bearing formations depends

on effective stress. The dependence is larger for consolidated sandstones. The detailed

discussions on pressure / stress dependent permeability are given in Chapter 2 along with

mathematical correlations derived by various authors. For the modeling of geomechanical

effects in this Chapter, the correlation presented by Jones and Owens (1980) was selected

and is given by Equation (2.8). According to their correlation data must be corrected for

1,000 psi effective pressure (stress) and stress factor, S, can be found by plotting

laboratory data on semi-log plot. Since for our project there is no laboratory data

available and stress dependent permeability curves are used as history matching

parameter, Equation (2.8) needs to be modified as follows.

where = Permeability at corresponding mean effective stress

= Permeability at initial mean effective stress

= Stress factor

= Mean effective stress

= Initial mean effective stress

Initial and corresponding mean effective stress is found using Equation (6.1). For

pressure dependent permeability condition stated in Equation (6.4) is used to generate

data that depends on fluid pressure by keeping the total mean stress constant throughout.

91
It is assumed that stresses in the reservoir are not changing due to change in pressure and

remain constant during production. Taking this assumption and using Equations (6.1) and

(6.5) pressure dependent permeability multipliers are calculated for different values of the

stress factor S.

6.3 HISTORY MATCHING USING PRESSURE DEPENDENT PERMEABILITY

6.3.1 Effects of Shape of Pressure Dependent Permeability Curves (Stress Factor, S)

Using Case 6.1 as a base case, few sub-cases were run to see the effects of the shape of

pressure dependent permeability multipliers, which is defined by the stress factor (S), on

gas rate and cumulative production. As fracture conductivity is taken as the realistic

fracture conductivity by considering all factors explained in Chapter 5, no pressure

dependent permeability multiplier is applied to blocks representing fractures. The initial

permeability is found at reference pressure of 2000 and 2200 psi for well A and B

respectively. Pressure dependent permeability curves of different shapes are generated

using range of values of stress factor by using Equation (6.5). The larger the value of S,

the larger is the permeability reduction and reservoir is more stress sensitive. It is

important to note that the same pressure dependent function is used in the whole reservoir

without distinguishing between high and low stress dependent regions. Summary of

assumed and calculated rock and fracture properties, which are discussed in more detail

in Chapter 4, for wells A and B are presented in Table 6.1 and 6.2 respectively. All the

properties that are discussed in base case (i.e., Case 6.1) will remain the same and in

92
addition pressure dependent multipliers with different stress factors are used. Stress

factors for sub Cases 6.3.1.1 to 6.3.1.4 are given in Table 6.3.

Table 6.3 Stress factor (S) of Cases 6.3.1.1 - 4 for well A and B

Property (Cases 6.3.1.1 - 4) Values – Well A Values – Well B

Case 6.3.1.1
Stress factor (Pi, psi = 2000 (A), 2200 (B)) 0.33
Case 6.3.1.2
Stress factor (Pi, psi = 2000 (A), 2200 (B)) 1.7
Case 6.3.1.3
Stress factor (Pi, psi = 2000 (A), 2200 (B)) 3.3
Case 6.3.1.4
Stress factor (Pi, psi = 2000 (A), 2200 (B)) 3.5 3.8

Pressure dependent permeability functions generated from stress factor values given in

Table 6.3 are presented in Figure 6.4. Figures 6.5 and 6.6 show the comparison of sub

cases in this current section with base case (i.e., Case 6.1) and field data followed by a

discussion on the results.

93
1.0

0.8

0.6 S=0.33 - Well A


S=1.7 - Well A
K(p)/Ki

S=3.3 - Well A
S=3.5 - well A
0.4
S=3.8 - Well A
S=0.33 - Well B
S=1.7 - Well B
0.2
S=3.3 - Well B
S=3.5 - Well B
S=3.8 - Well B
0.0
0 400 800 1200 1600 2000 2400
Fluid Pressure (psi)

Figure 6.4 Pressure dependent permeability curves of Cases 6.3.1.1 – 4 - well A & B

16 2800
Field Avg Gas Rate Gas Rate - C-6.1
Gas Rate - C-6.3.1.1 Gas Rate - C-6.3.1.2
Gas Rate - C-6.3.1.3 Gas Rate - C-6.3.1.4
Field Cumulative Gas Cumulative Gas - C-6.1
Cumulative Gas - C-6.3.1.1 Cumulative Gas - C-6.3.1.2
Cumulative Gas - C-6.3.1.3 Cumulative Gas - C-6.3.1.4

Cumulative Gas Prodcution (MMscf)


12 2100
Avg Daily Gas Rate (MMscfd)

8 1400

4 700

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.5 Comparison of field data and simulation results of Cases 6.3.1.1-4-well A
94
12 2800
Field Avg Gas Rate Gas Rate - C-6.1
Gas Rate - C-6.3.1.1 Gas Rate - C-6.3.1.2
Gas Rate - C-6.3.1.3 Gas Rate - C-6.3.1.4
Field Cumulative Gas Cumulative Gas - C-6.1
Cumulative Gas - C-6.3.1.1 Cumulative Gas - C-6.3.1.2
9 2100

Cumulative Gas Production (MMscf)


Cumulative Gas - C-6.3.1.3 Cumulative Gas - C-6.3.1.4
Avg Daily Gas Rate (MMscfd)

6 1400

3 700

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.6 Comparison of field data and simulation results of Cases 6.3.1.1-4-well B

The trend of the results shown in Figure 6.5 and 6.6 is similar except the last case. It is

showing very little deviation in very early time but the gap in results is starting to

increase after that. As the value of stress factor increases, permeability multiplier function

decreases leading to smaller gas rate and production. In all these cases permeability

multipliers are decreasing from 1 at initial pressure to some smaller value at minimum

bottom hole pressure constraint for well flowing pressure. The overall average reservoir

pressure is also declining resulting in smaller production due to corresponding overall

decrease of permeability. There is considerable difference in permeability multipliers of

well A and B at the same pressure; difference is larger at higher pressure. Now question

would be which curve to choose to define stress sensitivity of reservoir for this field?

Keeping in mind the unconventional nature of tight gas sands and high stress sensitivity

95
of tight reservoirs, Case 6.3.1.3 for well A and Case 6.3.1.4 for well B is chosen as the

base case for further investigation. Although both wells are producing from the same

formation and stress sensitivity will not be that different for each of them, different cases

have been selected as an initial guess.

6.3.2 Effects of Stimulated Reservoir Volume (SRV)

The effects of stimulated reservoir volume were already studied in detail in Chapter 5.

Case 5.3.6 presented in Table 5.5 gives very good match in early time. The purpose to

run a case in this section is to apply the SRV size assumed in Case 5.3.6 to base case,

which is selected for this section. It is important to mention that pressure dependent

function is applied both inside and outside SRV for the current simulation run. Table 6.4

and 6.5 summarized fracture and rock properties and also present comparison between

data sets for current case and base case for well A and B respectively.

Table 6.4 Fracture properties of Case 6.3.2 for well A

Values Values
Property – Well A
(Case 6.3.1.3) (Case 6.3.2)
Net pay thickness, ft 100
Fracture height, ft 100
Fracture width, inches 0.25
Fracture half length, ft 28
Fracture conductivity, mD-ft 39
Initial permeability multiplier 10 (0.15, 0.015)
Number of fractures, n 15
131
1301

96
Stress factor (Pi, psi = 2000 (A)) 3.3
SRV (one wing), ft3 120x3000x100 120x300x100

Table 6.5 Fracture properties of Case 6.3.2 for well B

Values Values
Property – Well B
(Case 6.3.1.4) (Case 6.3.2)
Net pay thickness, ft 100
Fracture height, ft 100
Fracture width, inches 0.125
Fracture half length, ft 63
Fracture conductivity, mD-ft 13
Initial permeability multiplier 12 (0.12, 0.012)
Number of fractures, n 15
55
543
Stress factor (Pi, psi = 2200 (B)) 3.8
SRV (one wing), ft3 140x3000x100 140x500x100

Comparison of results between Case 6.3.2 and corresponding base case for each well is

presented in Figures 6.7 and 6.8. The average fracture spacing is 180 ft, so to avoid any

fracture interference and to get maximum width created by microseismic events; the SRV

width should be smaller than average fracture spacing. Although the production of this

current case is smaller than in the base case, the shape of curve both at early time and late

resembles to field data. The SRV is chosen from Case 5.3.6 in which good match at early

time was achieved. Case 6.3.2 is selected as base case for both wells and will be used to

compare results.

97
16 1600
Field Avg Gas Rate Gas Rate - C-6.3.1.3
Gas Rate - C-6.3.2 Field Cumulative Gas
Cumulative Gas - C-6.3.1.3 Cumulative Gas - C-6.3.2

Cumulative Gas Prodcution (MMscf)


12 1200
Avg Daily Gas Rate (MMscfd)

8 800

4 400

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.7 Comparison of field data and simulation results of Case 6.3.2 - well A

12 1600
Field Avg Gas Rate Gas Rate - C-6.3.1.3
Gas Rate - C-6.3.2 Field Cumulative Gas
Cumulative Gas - C-6.3.1.3 Cumulative Gas - C-6.3.2

Cumulative Gas Production (MMscf)


9 1200
Avg Daily Gas Rate (MMscfd)

6 800

3 400

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.8 Comparison of field data and simulation results of Case 6.3.2 - well B

98
6.3.3 History Matching Using Pressure Dependent Permeability outside SRV

The effects of varying stress factor and stimulated reservoir volume was studied so far,

and both parameters had significant impacts on results. Until this point all the cases in

this section are run by applying pressure dependent permeability multiplier to the whole

reservoir. The secondary fractures in SRV are believed to be open initially but closing

rapidly in early time of production to some residual value. So permeability dependence

on pressure will be larger inside the SRV than outside of it. Cases in this section are run

by applying pressure dependent multipliers to the whole reservoir except inside SRV.

This may seem contradictory to what was stated in earlier chapters and sections. But

these cases were investigated and included here to highlight the importance of using

correct approach to get history match. High production both in early time and late time is

expected by removing pressure dependent permeability multipliers inside SRV so

constant (higher) permeability is used inside SRV.

Same fracture and rock properties given in Table 6.4 and 6.5 are used in this case. The

first case in this section, which is Case 6.3.3.1, is exactly similar to base case, i.e., Case

6.3.2 except there is no pressure dependent permeability inside the SRV in the former.

Figures 6.9 and 6.10 present result of field data, base case, and Case 6.3.3.1 for well A

and B respectively.

99
16 1200
Field Avg Gas Rate Gas Rate - C-6.3.2
Gas Rate - C-6.3.3.2 Field Cumulative Gas
Cumulative Gas - C-6.3.2 Cumulative Gas - C-6.3.3.1

Cumulative Gas Prodcution (MMscf)


12 900
Avg Daily Gas Rate (MMscfd)

8 600

4 300

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.9 Comparison of field data and simulation results of Case 6.3.3.1 - well A

12 1500
Field Avg Gas Rate Gas Rate - C-6.3.2
Gas Rate - C-6.3.3.1 Field Cumulative Gas
Cumulative Gas - C-6.3.2 Cumulative Gas - C-6.3.3.1
1200
9

Cumulative Gas Production (MMscf)


Avg Daily Gas Rate (MMscfd)

900

600

3
300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.10 Comparison of field data and simulation results of Case 6.3.3.1 - well B

100
As both wells are drilled in same formation, one would expect that the permeability

dependence on pressure or stress should be the same. Case 6.3.3.2 is run both for well A

and well B. Its purpose is to use same stress factor in both wells and improve results for

well B in late time because it was deviating at late time as shown in Figure 6.10. It is

important to recall that pressure dependent permeability multipliers (PDPM) are applied

to the whole reservoir excluding SRV. New set of parameters of fracture and rock for

well A and B are given in Table 6.6. The results of this case are presented in Figures 6.11

and 6.12 for well A and B respectively.

Table 6.6 Fracture properties of Case 6.3.3.2 for well A and B

Property (Case 6.3.3.2) Values – Well A Values – Well B


Net pay thickness, ft 100
Fracture height, ft 100
Fracture width, inches 0.25 0.125
Fracture half length, ft 28 63
Fracture conductivity, mD-ft 39 13
Initial permeability multiplier 12 (0.18, 0.018) 25 (0.25, 0.025)
109 27
1084 261
Stress factor outside SRV, S 3.8 3.8
SRV (one wing), ft3 120x300x100 140x400x100

101
16 1200
Field Avg Gas Rate Gas Rate - C-6.3.3.2

Field Cumulative Gas Cumulative Gas - C-6.3.3.2

Cumulative Gas Prodcution (MMscf)


12 900
Avg Daily Gas Rate (MMscfd)

8 600

4 300

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.11 History matched case using PDPM outside SRV - well A

12 1200
Field Avg Gas Rate Gas Rate - C-6.3.3.2

Field Cumulative Gas Cumulative Gas - C-6.3.3.2

Cumulative Gas Production (MMscf)


9 900
Avg Daily Gas Rate (MMscfd)

6 600

3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.12 History matched case using PDPM outside SRV - well B
102
Excluding permeability multipliers inside the SRV improved both gas rate and

cumulative production as shown in Figures 6.11 and 6.12. One interesting observation

from these figures is the similar production at very early time. Larger permeability inside

SRV than in the rest of the formation results in higher production in early time; at late

time production is controlled by region outside SRV so using permeability multiplier in

that region actually brings down production and results in good match.

Initial permeability multiplier of 12 and 25 is used to get match for well A and B

respectively while the microseismic events length (SRV length) was decreased by 100 ft

to get history match, which means that decrease in SRV length is compensated by

increasing initial permeability multiplier. Late time results are improved for well B by

implementing those changes, as can be confirmed by comparing Figures 6.10 and 6.12.

6.3.4 Effect of Native Reservoir Permeability – Well B

Simulation results shown in Figure 6.11 of Case 6.3.3.2 for well B are 3.80% larger than

field data at the end. This seems a reasonable match for current field data but difference

will grow if the simulation is run for longer time. So far all history matching parameters

were investigated but the model still needs further improvements and tunings. Since well

B is deeper than well A, it could be possible that native reservoir permeability initially

selected for well B is different. For this purpose few cases were run and effects of native

reservoir permeability on production were investigated and evaluated. The new reservoir

permeability is decreased by factor of 2, which gives 0.005 mD. Case 6.3.3.2 is selected

103
as base case for these simulation runs. The only difference between Case 6.3.3.2 and

Case 6.3.4.1 is use of smaller (by factor of 2) native reservoir permeability.

Case 6.3.4.2 is run by taking Case 6.3.4.1 as base case in which microseismic length and

initial permeability multiplier is modified to get history match. Fracture properties and

summary of rock properties used in all of the above discussed cases are presented in

Table 6.7. Figure 6.13 represents comparison of results between base case and sub cases

from this section. Figure 6.14 shows history matched case result with field data using

pressure dependent permeability outside SRV and modified native reservoir permeability.

Starting with comparison of base case and Case 6.3.4.1 using smaller permeability will

decrease production, which is obvious and it is shown in Figure 6.13. The entire

production curve shifts downward, which is then pushed upward by increasing

microseismic length from 400 to 500 ft and using larger permeability multiplier for Case

6.3.4.2. Initial permeability capture early time production as well as late time but major

part of late time production is increased by increasing microseismic length.

Table 6.7 Fracture properties of Cases 6.3.4.1 - 2 for well B

Values Values Values


Property – Well B
(Case 6.3.3.2) (Case 6.3.4.1) (Case 6.3.4.2)
Fracture height, ft 100
Fracture width, inches 0.125
Fracture half length, ft 63
Stress factor outside SRV, S 3.8
Fracture conductivity, mD-ft 13
25 25 50
Initial permeability multiplier
(0.25, 0.025) (0.125, 0.0125) (0.25, 0.025)
Reservoir permeability, mD 0.01, 0.001 0.005, 0.0005

104
27 53 27
261 521 261
SRV (one wing), ft3 140x400x100 140x400x100 140x500x100

12 1200
Field Avg Gas Rate Gas Rate - C-6.3.3.2
Gas Rate - C-6.3.4.1 Gas Rate - C-6.3.4.2
Field Cumulative Gas Cumulative Gas - C-6.3.3.2
Cumulative Gas - C-6.3.4.1 Cumulative Gas - C-6.3.4.2
9 900

Cumulative Gas Production (MMscf)


Avg Daily Gas Rate (MMscfd)

6 600

3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.13 Comparison of field data and simulation results of Case 6.3.4.1-2-well B

The results for history matched case model by decreasing native reservoir permeability

given in Figure 6.14 are much improved compared to previous results shown in Figure

6.12.

From all these simulation runs and exercises, it is deduced that similar results can be

achieved by varying various parameters but question remains about the validity of the

various scenarios that were used to force a match which do not account for all physics

actually happening in reservoir. History matching so far excluded pressure dependent

105
permeability in SRV. Pressure dependent permeability inside SRV cannot be ignored

and will be considered next.

12 1200
Field Avg Gas Rate Gas Rate - C-6.3.4.2

Field Cumulative Gas Cumulative Gas - C-6.3.4.2

9 900

Cumulative Gas Production (MMscf)


Avg Daily Gas Rate (MMscfd)

6 600

3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.14 History matched case using PDPM outside SRV and modified reservoir

permeability - well B

6.3.5 History Matching Using Pressure Dependent Permeability both Inside and

Outside SRV

In section 6.3.4 pressure dependent permeability was used outside SRV only and the

purpose of simulation runs in this section is to get history match with field data using

pressure dependent permeability both inside and outside SRV. Current case represent

more realistic approach than cases run in section 6.3.4 as smaller permeability multipliers

are applied inside SRV and larger outside SRV.

106
Fracture and rock properties for Case 6.3.5 and previously history matched case, i.e.,

Case 6.3.3.2 and 6.3.4.2 for well A and B respectively, are summarized in Table 6.8. The

pressure dependent curves for history matched model case both for well A and B with

modified stress factor are presented in Figure 6.15. The results of current case and field

data are shown in Figures 6.16 and 6.17 for well A and B respectively.

Table 6.8 Fracture properties of Case 6.3.5 both for well A and B

Values Values
Property – Well A
(Case 6.3.3.2) (Case 6.3.5)
Fracture width, inches 0.25
Fracture half length, ft 28
Fracture conductivity, mD-ft 39
SRV (one wing), ft3 120x300x100
Stress factor Inside SRV, S No Multipliers 3.9
Stress factor outside SRV, S 3.8 2.7
109 27
1084 261
Initial permeability multiplier 12 (0.18, 0.018) 50 (0.75, 0.075)

Values Values
Property – Well B
(Case 6.3.4.2) (Case 6.3.5)
Fracture width, inches 0.125
Fracture half length, ft 63
Fracture conductivity, mD-ft 13
Reservoir permeability, mD 0.01, 0.001 0.005, 0.0005
SRV (one wing), ft3 140x400x100 140x600x100
Stress factor Inside SRV, S No Multipliers 3.9
Stress factor outside SRV, S 3.8 2.7
27 14
261 131
Initial permeability multiplier 25 (0.25, 0.025) 100 (0.5, 0.05)

107
Mean Effective Stress (Psia)
3500 4000 4500 5000 5500 6000
1
S=2.7 - Well A S=2.7 - Well B
S=3.9 - Well A S=3.9 - Well B
S=2.7 - Well A S=2.7 - Well B
0.8
S=3.9 - Well A S=3.9 - Well B

0.6
K(p)/Ki

0.4

0.2

0
0 300 600 900 1200 1500 1800 2100 2400
Pressure (Psia)

Figure 6.15 Pressure and effective stress dependent permeability curves at stress
factor of 2.7 and 3.9 - well A and B
16 1200
Field Avg Gas Rate Gas Rate - C-6.3.5

Field Cumulative Gas Cumulative Gas - C-6.3.5

12 900

Cumulative Gas Prodcution (MMscf)


Avg Daily Gas Rate (MMscfd)

8 600

4 300

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.16 History matched case using PDPM both inside and outside SRV- well A

108
12 1200
Field Avg Gas Rate Gas Rate - C-6.3.5

Field Cumulative Gas Cumulative Gas - C-6.3.5

9 900

Cumulative Gas Production (MMscf)


Avg Daily Gas Rate (MMscfd)

6 600

3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.17 History matched case using PDPM both inside and outside SRV - well B

The results shown in Figures 6.16 and 6.17 do not exactly matches at mid time of

production for both wells. The results obtained from current cases are not as good as

shown in Figures 6.11 for well A and 6.14 for well B but current cases represent more

realistic approach and should be closer to reality than previous cases. It is important to

highlight that in all preceding sections pressure dependent permeability multipliers are

calculated using mean effective stress where total mean stress is kept constant and only

fluid pressure is varied, concept defined by Equation (6.4). This implies that stresses near

wellbore and in reservoir are not changing during production and remain constant

throughout and effectives stress changes as fluid pressure changes.

109
6.4 APPROXIMATION OF GEOMECHANICAL EFFECTS IN

CONVENTIONAL RESERVOIR SIMULATOR USING UNIAXIAL

DEFORMATION THEORY

Coupled geomechanical modeling in reservoir simulation is the correct way to represent

effective stress dependent porosity and permeability, which calculates stresses of porous

media using concept of deformations. Complexity of models and cost of geomechanical

modeling make it very difficult to run it all the time. Geomechanical effects can be

modeled in conventional reservoir simulators (uncoupled model) without running

geomechanical model by reasonable approximations. Most commercial reservoir flow

simulators offer options to include pressure dependent porosity and permeability. Since

rock data is measured in laboratory as function of effective stress, there is always

confusion how to convert lab data that only depends on pressure for conventional

reservoir simulation. Settari et al. (2005) presented approximations for porosity (rock

compressibility) coupling that is mainly dependent on type of deformations and

permeability coupling by introducing formulations both for global and local stress

changes caused by pressure depletion.

Although it is true that there is more complex deformation near wellbore or fractures,

uniaxial deformation theory serves as a reasonable approximation when reservoir is

producing at uniform average pressure. Then the stress changes can be calculated as a

function of pressure depletion.

Hook’s law for homogenous and isotropic solid materials in three dimensions is (Fjær et

al., 2008)

110
where = Elastic modulus, it is the measure of resistance of elastic deformation

= Strain in corresponding direction where i = x, y, z

= Stress in corresponding direction

= Poisson’s ratio, it is ratio of lateral expansion to longitudinal contraction

= Modulus of rigidity or shear modulus, it the measure of resistance of shear

deformation =

Porous media also contains fluid in its pores, which has a large impact on the physical

and mechanical behavior of rock, which will be considered in next phase. The first linear

theory of isothermal Poroelasticity was developed by Maurice A. Biot (1941), a Belgian-

American physicist. According to Biot, the presence of the pore fluid adds extra terms to

the strain energy of the material. He assumed that pore fluid cannot produce any shearing

strain and effect of pore fluid must be the same on all components of strain. Hence

Equation (6.6) for isotropic and isothermal linearly poroelastic material using positive

compressive stress and pressure becomes (Fjær et al., 2008)

111
where = Fluid pressure

= Poroelastic constant

For linearly isotropic material the stress–strain behavior of material is fully described

only by 2 constants i.e. Young’s Modulus (E) and Poisson’s ratio (υ). The others

constants can be calculated from these two constants. The poroelastic constant is given by

where = Bulk drained compressibility

= Grain (solid rock or rock matrix) compressibility

= Bulk modulus of framework or frame modulus =

= Solid grain modulus =

Bulk modulus, which is the measure of resistance of sample rock to hydrostatic

compression, can be represented in following form.

where = Volumetric strain =

= lame’s 1st parameter =

Then last term in Equation (6.7) is

Where = Biot’s constant =

Putting Equation (6.10) into (6.7) it becomes

112
The following assumptions are now made to derive relations for stress changes in

poroelastic media based on uniaxial deformation theory. The properties of rock in each

direction are constant, i.e., isotropic material.

 The rock material behaves as linearly elastic material.

 Entire reservoir undergoes uniform pressure change.

 Poisson ratio is same in all direction due to isotropy of material.

 Reservoir is deforming uniaxially in vertical direction.

 Vertical stress is constant, i.e., change in vertical stress is zero ( or = 0).

 No lateral strains, i.e., = =0

 No shear strain, i.e., =0

 No arching effects.

Applying it to Equation (6.11) and neglecting shear stress it is reduced to

Solving for first two terms in equation (6.12) and simplifying terms

Where = Poroelastic constant =

113
This expression means that changes in stress in x-direction and y-direction do not depend

on the modulus of rock but only on Poisson’s ratio, Biot’s constant and change in fluid

pressure. Poroelastic constant varies between 0 and . The maximum value of Poisson

ratio is normally 0.5, fluids, and minimum is zero, brittle rocks. For the same type of

rock, changes in stress depend only on the change in pore pressure of fluid.

Equation (6.13) in form of in-situ minimum and maximum horizontal stress can be

written as

If total mean stress is kept constant shown in Equation (6.15) then permeability becomes

function of pressure defined by Equation (6.4). The data from lab can be used directly by

modifying it. As changes in mean effective stress are equal to change in fluid pressure

then equation simplifies to Equation (6.16).

Lets instead assume that each grid block deform uniaxially independent of all other grid

blocks. If fluid pressure in particular grid block changes from to then change in

pressure is

Effective mean stress when total mean stress is changing with pressure at any particular

pressure is

Change in total mean stress is equal to

114
Putting Equation (6.14) into (6.2)

Substituting Equation (6.19) into (6.18) results in

Mean effective stress, which of function of both total mean stress and fluid pressure

defined by Equation (6.18) is given as

Equation (6.21) represents more rigorous approach as compared to Equation (6.16) by

taking into account poroelastic effects using uniaxial deformation theory. If permeability

is a function of total mean stress and fluid pressure then permeability multipliers

corrected for geomechanical effects can be calculated using Equation (6.22). Equation

(6.15) is a function of fluid pressure only and total mean stress is constant and Equation

(6.22) is function of both fluid pressure and varying total mean stress. The mean effective

stress for variable total mean stress calculated by Equation (6.22) will be smaller by

factor of than for constant total mean stress calculated by Equation (6.15).

115
6.5 HISTORY MATCHING USING PRESSURE DEPENDENT PERMEABILITY

BOTH INSIDE AND OUTSIDE SRV – COUPLED PRODUCTION

GEOMECHANICAL MODELS

To emphasize the importance of use of proper data set for any type of simulation, some

cases are run. Case 6.3.5 both for well A and B is considered as base for current case.

Fully coupled geomechanical models are run using effective stress dependent

permeability shown in Figure 6.14. The difference between Case 6.3.5 and current case is

that former is uncoupled reservoir flow model while the latter is a fully coupled

geomechanical model. Permeability multipliers are used in form of table in reservoir

simulators. In uncoupled model dependent variable is fluid pressure while in coupled

geomechanical model dependent variable is effective mean stress. In current case

permeability generated for uncoupled model is used in coupled geomechanical simulation

without correcting data set for poroelastic effects. Same permeability multipliers are

applied in input file for coupled simulation without modifying data set and table of mean

effective stress and multipliers are generated by changing fluid pressure given by

Equation (6.15).

Fracture and rock properties of Case 6.3.5, i.e., base case are given in Table 6.8 both for

well A and B are used here for current case, i.e., Case 6.5. The mean effective stress

dependent permeability curves shown in Figure 6.15 as dashed lines are used in coupled

geomechanical models for both wells. The results of Case 6.5 and Case 6.3.5 for both

wells and their comparison with field data are presented in Figures 6.18 and 6.19.

116
16 1600
Field Avg Gas Rate Gas Rate-C-6.3.5-Uncoupled-CMS*
Gas Rate-C-6.5-Coupled-CMS Field Cumulative Gas
Cumulative Gas-C-6.3.5-Uncoupled-CMS Cumulative Gas-C-6.5-Coupled-CMS
* CMS - Constant total mean stress

Cumulative Gas Prodcution (MMscf)


12 Stress 1200
Avg Daily Gas Rate (MMscfd)

8 800

4 400

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.18 Comparison of field data and simulation results of uncoupled and
coupled cases: Constant mean total stress - well A
12 1600
Field Avg Gas Rate Gas Rate-C-6.3.5-Uncoupled-CMS*
Gas Rate-C-6.5-Coupled-CMS Field Cumulative Gas
Cumulative Gas-C-6.3.5-Uncoupled-CMS Cumulative Gas-C-6.5-Coupled-CMS
* CMS - Constant total mean stress

Cumulative Gas Production (MMscf)


9 Stress 1200
Avg Daily Gas Rate (MMscfd)

6 800

3 400

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.19 Comparison of field data and simulation results of uncoupled and
coupled cases: Constant mean total stress - well B

117
Running coupled model with the same data set results in higher production for both wells

as shown in Figures 6.18 and 6.19. Change in pressure given in Equation (6.17) is always

negative in depletion (production) case so mean effective stress calculated by Equation

(6.22) will always be smaller than calculated by Equation (6.15) for production case and

larger for injection case. In coupled reservoir simulation, fluid pressure serves an input to

geomechanical model which then calculates total stresses as function of change in

pressure, Poisson’s ratio and Biot’s constant. The reservoir model then takes input as

effective stress from geomechanical model and computes permeability multipliers from

that corresponding effective stress. The above discussion can be summarized in the

following flow chart.

Figure 6.20 Coupling of reservoir and geomechanical simulator – An overview

The plot of effective mean stress calculated by Equation (6.22) for variable total mean

stress, and by Equation (6.15) for constant total mean stress is given in Figure 6.21.

118
5688
Constant Mean Total Stress - Eq (6.15) - Well A
Variable Mean Total Stress - Eq (6.22) - Well A

5188 σm’ = 5188 @ P = 500 psia


Mean Effective Stress (Psi)

4688

σm’ = 4331 @ P = 500 psia

4188

3688
0 500 1000 1500 2000
Pressure (Psi)

Figure 6.21 Plot of mean effective stress and fluid pressure – well A

The effective stress is smaller for variable mean total stress case than constant mean total

stress case due to poroelastic effects. It means that total stresses are smaller in porous

media due to poroelasticity. Plotting permeability multipliers calculated using Equation

(6.5) both for constant and variable mean total stress cases against mean effective stress

and fluid pressure will further clarify why the results in coupled simulation case are

higher than in the uncoupled case. Total mean stresses calculated by geomechanical

simulator are not constant but always changing due to poroelasticity, which results in

smaller values at corresponding pressure. So permeability multiplier will be larger for

coupled case and hence higher gas rate and cumulative production. For the same value of

pressure, permeability multiplier for variable mean total stress case is always larger than

for the constant mean total stress case due to smaller values of mean effective stress as

shown in Figure 6.22.

119
1
S=2.7 - Well A - Constant Total Mean Stress
S=2.7 - Well A - Variable Total Mean Stress

0.8 S=3.9 - Well A - Constant Total Mean Stress


S=3.9 - Well A - Variable Total Mean Stress

0.6

K(p)/Ki

0.534 @ σm = 4331

0.4

0.368 @ σm = 4331


0.216 @ σm = 5188
0.2


0.075 @ σm = 5188
0
3688 4188 4688 5188 5688
Mean Effective Stress (Psi)

Figure 6.22 Permeability multipliers and mean effective stress for constant and

variable mean total stress cases – well A

Although all the discussion and examples are for well A, same conclusions apply to well

B.

6.6 HISTORY MATCHING USING PRESSURE DEPENDENT PERMEABILITY

BOTH INSIDE AND OUTSIDE SRV – COUPLED PRODUCTION

GEOMECHANICAL MODELS – VARIABLE MEAN TOTAL STRESS

CASES

Based on discussion in previous section, it is necessary to correct the data set in order to

get realistic history match and forecast. Permeability multipliers computed using variable

total mean stress will be the correct way of modeling geomechanical effects in reservoir

120
simulation. For this purpose first of all permeability multipliers are recalculated using

Equation (6.22) and compared with previous multipliers used in section 6.3.5. The

modified set of permeability multipliers with larger stress factor for current case and

previous cases both for well A and B are presented in Figures 6.23 and 6.24.

Mean Effective tress (Psia)


3688 4188 4688 5188 5688
1
S=2.7-CMS* S=6.0-VMS** S=3.9-CMS S=8.4-VMS
S=2.7-CMS S=6.0-VMS S=3.9-CMS S=8.4-VMS

0.8 * CMS - Constant total mean stress


** VMS-Variable total mean stress

Outside SRV
0.6
K(p)/Ki

Inside SRV
0.4

0.2

0
0 500 1000 1500 2000
Pressure (Psi)

Figure 6.23 Pressure and effective stress dependent permeability curves at stress

factor of 2.7, 3.9, 6 and 8.4 - well A

Simulation is run using modified set of data for both wells using uncoupled and fully

coupled simulators. Case 6.6 is the case for both wells in which permeability multipliers

are calculated by considering poroelastic effects. Fracture and rock properties used in

current case and their comparison with previously history matched cases are presented in

Table 6.9. Simulation results are presented in Figures 6.25 and 6.26 for well A and B

respectively.

121
Mean Effective Stress (Psi)
3553 4053 4553 5053 5553 6053
1
S=2.7-CMS S=6.0-VMS S=3.9-CMS S=8.4-VMS
S=2.7-CMS S=6.0-VMS S=3.9-CMS S=8.4-VMS

0.8 * CMS - Constant total mean stress


** VMS-Variable total mean stress
Outside SRV
0.6 Inside SRV
K(p)/Ki

0.4

0.2

0
0 500 1000 1500 2000 2500
Pressure (Psi)

Figure 6.24 Pressure and effective stress dependent permeability curves at stress

factor of 2.7, 3.9, 6 and 8.4 - well B

Table 6.9 Fracture properties of Cases 6.6 both for well A and B

Values Values
Property – Well A
(Case 6.3.5) (Case 6.6)
Fracture height, ft 100
Fracture width, inches 0.25
Fracture half length, ft 28
Fracture conductivity, mD-ft 39
Number of fractures, n 15
SRV (one wing), ft3 120x300x100
27
261
Initial permeability multiplier 50 (0.75, 0.075)
Stress factor Inside SRV, S 3.9 8.4
Stress factor outside SRV, S 2.7 6.0

122
Values Values
Property – Well B
(Case 6.3.4.5) (Case 6.6)
Fracture height, ft 100
Fracture width, inches 0.125
Fracture half length, ft 63
Fracture conductivity, mD-ft 13
Number of fractures, n 15
Reservoir permeability, mD 0.005, 0.0005
SRV (one wing), ft3 140x600x100
14
131
Initial permeability multiplier 100 (0.5, 0.05)
Stress factor Inside SRV, S 3.9 8.4
Stress factor outside SRV, S 2.7 6.0

16 1200

12 900

Cumulative Gas Prodcution (MMscf)


Avg Daily Gas Rate (MMscfd)

8 600

Field Avg Gas Rate Gas Rate-C-6.6-Uncoupled-VMS*

Gas Rate-C-6.6-Coupled-VMS Field Cumulative Gas

Cumulative Gas-C-6.6-Uncoupled-VMS Cumulative Gas-C-6.6-Coupled-VMS


4 300
* VMS - Variable total mean stress

0 0
12-Jun-06 14-Dec-06 18-Jun-07 20-Dec-07 23-Jun-08 26-Dec-08 29-Jun-09 1-Jan-10

Figure 6.25 Comparison of field data and simulation results of uncoupled and

coupled cases: Variable mean total stress - well A

123
12 1200

9 900

Cumulative Gas Production (MMscf)


Avg Daily Gas Rate (MMscfd)

6 600
Field Avg Gas Rate Gas Rate-C-6.6-Uncoupled-VMS*

Gas Rate-C-6.6-Coupled-VMS Field Cumulative Gas

Cumulative Gas-C-6.6-Uncoupled-VMS Cumulative Gas-C-6.6-Coupled-VMS

* VMS - Variable total mean stress


3 300

0 0
16-Aug-06 8-Feb-07 3-Aug-07 26-Jan-08 21-Jul-08 13-Jan-09 8-Jul-09 1-Jan-10

Figure 6.26 Comparison of field data and simulation results of uncoupled and

coupled cases: Variable mean total stress - well B

Results of both uncoupled and coupled geomechanical production models are remarkably

similar as shown in Figures 6.25 and 6.26. This shows that proper care should be taken

while running coupled geomechanical model with a data set from an uncoupled model.

Stresses are changing constantly so constant total mean stress assumption is not an

accurate assumption; it must be replaced with variable total mean stress. Poroelastic

effects can be modeled in conventional reservoir simulator by assuming uniaxial

reservoir deformation. Matrix permeability enhances during fracturing, which is

represented here by initial permeability multiplier (IPM). IPM of 50 and 100 is used for

124
well A and B respectively to get history match. These numbers are consistent with those

reported by Settari et al. (2002b) and Settari et al. (2009).

6.7 SUMMARY

 Stimulated reservoir region is much larger than propped fracture length.

 The dimensions of Stimulated reservoir volume (SRV) are different for x-linked

and slick water fracturing wells – shorter for x-linked fracture job.

 One possible way to get history match is the use of a constant permeability

multipliers inside SRV and pressure dependent permeability (K (p)) outside SRV.

 The two wells seem to have somewhat different permeability.

 Accurate way to get history match is to use pressure dependent permeability

multipliers both inside and outside SRV.

 Permeability of the region inside SRV has larger pressure dependence than region

outside SRV due to presence of complex network of microfractures.

 Moving from uncoupled to fully coupled geomechanical modeling without

correcting data sets for poroelastic effects will result in higher production. Data

sets must be correct for geomechanical aspects.

 Constant mean total stress assumption is not realistic approach as horizontal

stresses are decreasing with pressure depletion. Such assumption will always

result in overestimated production.

125
 New method based on uniaxial deformation theory is proposed to approximate

poroelastic effects in the reservoir and it predicts variation of mean total stress

depending on Poisson’s ratio, Biot’s constant and fluid pressure.

 The method proves to be a good approximation to capture poroelastic effects and

can be used in conventional reservoir simulation without running fully coupled

geomechanical simulation.

 Good history match is achieved in uncoupled simulation when data set is

corrected for poroelastic effects and results are confirmed by comparing it with

results from coupled geomechanical simulation.

126
CHAPTER SEVEN: HISTORY MATCHING OF FIELD INJECTION

PRESSURE USING UNCOUPLED AND FULLY COUPLED

GEOMECHANICAL INJECTION MODELS

It was evident in prior sections that history matching of field production data did not give

a unique set of parameters. Models with a different set of parameters can give almost the

same results and different sub-sets of parameters with different set of assumptions can

predict the same results. Injection models can be used to verify the number of parameters

involved in production history matching by matching well injection pressure during

treatment with simulation results. Field bottomhole injection pressure (BHIP) both for

well A and B is given in Figure 3.5. Well B which is deeper than A has a higher

bottomhole injection pressure than well A. Breakdown pressure, maximum required to

initiate fracture in formation, is also higher for well B.

This chapter represents simplified modeling of fracture propagation (dynamic fractures),

and history matching of field injection pressure using uncoupled and fully coupled

geomechanical injection models. The sensitivity study of various parameters such as

permeability enhancement/reduction factor, length of fracture propagation, and Biot’s

constant is performed in this chapter. Limitations of the uncoupled and coupled

geomechanical injection models are also discussed in this chapter. Figure 7.1 represents

flow chart of simulation runs performed in this chapter. Every previous case serves as

base case for next simulation run.

127
History Matching of Field Injection Pressure using
Uncoupled and Coupled Geomechanical Injection Models

Uncoupled Geomechanical Injection Coupled Geomechanical Injection


Models Models

Effects of Permeability Reduction Effects of Limiting Length of Fracture


Factor (Rfa) – Cases – 7.3.1.1 – 3 Propagation – Cases – 7.4.1.1 – 4

Effects of Limiting Length of Fracture Effects of Biot’s Constant (α) –


Propagation – Cases – 7.3.2.1 – 2 Cases – 7.4.2.1 – 2

Failure Predictions
Mohr – Coulomb Failure Criterion (Shear Fracturing)

Figure 7.1 Flow chart of simulation cases – Chapter 7

7.1 MODELING OF FRACTURE PROPAGATION

Fracture propagation modeling is an important part of physics and must be considered in

injection modeling of wells. Classical modeling of fracture geometry is well established

and documented in literature of rock mechanics and stimulation (Economides and Nolte,

2000). Coupling of fracture propagation (fracture dynamics) and fluid flow is

computationally very expensive (Ji et al., 2004b, 2009). The modeling of “complex”

fracturing discussed in Chapter 2 (Weng et al., 2011) is also expensive. However, proper

representation of dynamic propagation in which the fracture is directly coupled into a

reservoir simulator is important for many applications. Settari (1990, 2002a) has shown

128
that the fully coupled treatment of fracture mechanics, reservoir modeling and

geomechanics is important for better understanding of the unconventional fracturing

applications (waterflooding at fracturing condition, etc.) as well as for tight gas fracturing

treatments such as waterfracs and will be described briefly here. Such coupling, described

next, can be achieved in a simplified fashion which makes the modeling computationally

efficient.

7.2 THEORY OF SIMPLIFIED FRACTURE MODEL

The method is based on modifications of transmissibility in the reservoir flow model as

discussed in Chapter 4. For injection scenarios transmissibilities are modified

dynamically. To model dynamic fracturing process, a transmissibility multiplier function

is assigned to a line (or plane) of grid blocks assumed for fracture propagation extending

around the well. The multiplier function is a table that can be derived from simple 2-D

analytical fracture models which approximates the actual fracture. In an uncoupled

modeling, transmissibility multipliers are a function of fracture injection pressure, while

in a coupled (geomechanical) system they are a function of a minimum effective stress.

The multipliers are calculated based on the estimation of a 2D crack opening in a cross-

section by Equation (7.2) (Sneddon, 1969), then calculating the fracture permeability by

Equation (7.1) as a function of the net pressure in the fracture, and finally calculating the

multiplier as described in Equation (7.3).

129
Where = Fracture permeability

= constant, permeability reduction factor

= Fracture width

= Fracture treatment pressure (fluid pressure)

= Fracture opening or closing pressure

= Fracture half height

= Poisson’s ratio

= Elastic modulus

= Transmissibility multiplier

= Matrix permeability

= grid width

Hf in Equation (7.2) is the fracture half-height based on the 2-D Perkins-Kern geometry

assumption of vertical fracture with smooth closure at the top and bottom (Perkins and

Kern, 1961). Permeability reduction factor ( ) is a constant and it is one of the history

matching parameters. Fracture opening or closing pressure ( ) is normally taken as

initial minimum horizontal stress acting perpendicular to fracture face. may be

actually greater (or lower) than the initial minimum horizontal stress due to poroelastic

and thermo elastic effects. For uncoupled modeling it can be assumed that the stress

changes have stabilized and fracture opening or closing pressure is equal to the average

130
(adjusted) at minimum horizontal stress during the treatment. So transmissibility in

Equation (7.3) is function of fracture treatment pressure only for uncoupled modeling.

Where = Initial minimum horizontal stress (or adjusted value)

7.3 HISTORY MATCHING OF FIELD INJECTION PRESSURE USING

UNCOUPLED INJECTION MODELS

Hydraulic fracturing alters the initial in-situ stresses particularly around wellbore and

fractures. Although estimation of stress changes due to injection pressure using uniaxial

deformation theory is not accurate representation, it serves as a reasonable approximation

for this project. The Concept used in Chapter 6 for approximation of geomechanical

effects in an uncoupled model and equations developed for production cases, i.e.,

Equation (6.15) and (6.22) can be used also for injection modeling. For injection cases

change in pressure is always positive given by Equation (6.15). So effective mean stress

calculated by Equation (6.22) considering poroelastic effects is always larger than

calculated by Equation (6.15), which ignores poroelasticity. This means that

pressurization during fracturing increases stresses and stresses should be modified

accordingly. Changes in total stresses in x, and y – direction are calculated by the same

method as discussed for production modeling. History matching field injection pressure

using uncoupled injection models is different from coupled simulation because there is no

stress solution in uncoupled model so any changes in stresses due to poroelastic effects

131
must be incorporated manually in reservoir simulator for either permeability multipliers

or transmissibility calculation. Modified stresses are used in uncoupled model assuming

that hydraulic fracture increases in-situ stresses near wellbore and around fractures due to

poroelasticity and permeability enhancement in that region must be predicted by using

these modified stresses. For this purpose simple approach is used; minimum horizontal

stress was estimated from field injection pressure data and values are compared with

initial in-situ stress values calculated by data given in Table 3.3. The factor by which

both values are different is taken as multiplying factor to calculate modified in-situ stress

(stresses after fracturing). Initial and modified in-situ stresses in each corresponding

direction for wells A and B are presented in Table 7.1.

Fracture propagation in reservoir simulator is represented by modifying transmissibility

of fracture plane against corresponding injection pressure. Transmissibilities are

calculated using Equation (7.3) which is used in a form of table vs. injection pressure in

reservoir simulator blocks and fracture is allowed to grow without restricting its

geometry, i.e., half length and width. The data used to calculate transmissibility for wells

A and B is presented in Table 7.2.

Table 7.1 Initial and modified stresses for injection cases – Well A and B

Well A Well B
Initial Modified Initial Modified
value value value value
Vertical Stress, psi ( ) 6572 8400 6646 9090
Maximum Horizontal Stress, psi ( ) 5989 7800 6056 8500
Minimum Horizontal Stress, psi ( ) 4505 6300 4556 7000

132
Table 7.2 Data for transmissibility calculations in injection cases – Well A and B

Property Well A Well B


Fracture half height, ft 50
Pressure constraint, psia 10,000
Poison ratio, fraction 0.125
Grid block width, ft 2
Minimum horizontal stress, psia 6300 7000
Elastic modulus, psia 7.99E6 7.99E6
Matrix permeability, mD (Vertical) 0.015 (0.0015) 0.005 (0.0005)

In Chapter 6, it was shown with field examples that pressure / stress dependent

permeability increases due to poroelastic effects. In section 6.6 of Chapter 6 stress

dependent matrix permeability multipliers outside SRV are defined by stress factor of 6

for history matched cases. Permeability multipliers for uncoupled injection cases are

calculated using Equation (6.5) while effective mean stress is determined from Equation

(6.22). Modified in-situ stresses given in Table 7.1 are used for permeability multipliers

calculations. Calculated permeability multipliers both for wells A and B are presented in

Figure 7.2. Injection model was setup in reservoir simulator and run for given injection

period by setting maximum bottomhole injection pressure of 10,000 psi.

133
Mean Effective Stress (psi)
2000 3000 4000 5000 6000
50
Injection - Well A - S = 6
Injection - Well B - S = 6
Injection - Well A - S = 6
40 Injection - Well B - S = 6

30
K(p)/Ki

20

10

0
2000 4000 6000 8000 10000
Bottomhole Injection Pressure (psi)

Figure 7.2 Pressure and effective stress dependent permeability curves at stress
factor of 6 (uncoupled injection cases) - well A and B

7.3.1 Effects of Permeability Reduction Factor (

Permeability reduction factor plays an important role to calculate transmissibility of grid

blocks in fracture plane. Fracture permeability given by Equation (7.1) becomes an

intrinsic permeability of smooth open fracture if reduction factor is taken as 1. Fracture

permeability in reality is much smaller than this value due to tortuosity, asperities

interlocking, rock chipping at fracture face, unequal and rough surface of rock faces, and

fracture degradation. All these factors tend to decrease fracture permeability by order of

magnitude and are represented by the permeability reduction factor. So fracture

permeability calculated by Equation (7.1) with a factor < 1 represents more realistic

situation. Three injection cases are run to match field injection pressure with simulation

134
injection pressure by varying only the fracture permeability reduction factor. Fracture

propagation was allowed in y-direction without restricting or confining its half length.

Transmissibility given by Equation (7.3) is dependent on permeability reduction factor so

transmissibility values are calculated using data set given in Table 7.2 for each case with

the corresponding reduction factor. Matrix permeability multipliers are same in all

direction, i.e., x, y and z, and given in Figure 7.2 for these cases, and are applied in all

grid blocks except blocks in fracture plane because fracture transmissibility multipliers

are used for those blocks. The purpose of running these cases is to see the effects of

permeability reduction factor Rfa on injection pressure. Reduction factors for each case

run for wells A and B are given in Table 7.3.

Table 7.3 Permeability reduction factor of Cases 7.3.1.1-4 both for well A and B

Property Cases Well A Well B


Case 7.3.1.1 0.1
Permeability
Case 7.3.1.2 0.01
Reduction Factor
Case 7.3.1.3 0.001

Transmissibility multipliers are calculated for the above three cases using corresponding

values of permeability reduction factor by Equation (7.3) and are plotted in Figure 7.3;

similar graph can be constructed for well B. Transmissibility in x-direction is equal to 1.0

as fracture is growing only in y-z plane. Simulation results of bottomhole injection

pressure for all the three cases and field injection pressure are presented in Figure 7.4 and

7.5 for well A and B respectively.

135
1.0E+12

1.0E+09
Transmissibility Multipliers

1.0E+06
TmultY-Rfa=0.1
TmultZ-Rfa=0.1
TmultY-Rfa=0.001
1.0E+03 TmultZ-Rfa=0.001
TmultY-Rfa=0.01
TmulyZ-Rfa=0.01

1.0E+00
6000 7000 8000 9000 10000
Bottomhole Injection Pressure (psi)

Figure 7.3 Transmissibility multipliers – uncoupled injection cases - well A

Results shown in Figures 7.4 and 7.5 show that decreasing reduction factor pushes the

injection pressure upward. Smaller reduction factor means smaller transmissbility

mutlipliers and hence frcature propagation becomes more difficult due to larger pressure

drop down in the fracture. Smaller matrix permeability and fracture propagation hinder

fluid movement and pressure builds up. The injection pressure curves from simulation

show almost constant pressure trend throughout but slowly increasing with time.

136
8000
BHIP - Field - Well A
BHIP - C-7.3.1.1
BHIP - C-7.3.1.2
Bottomhole Injection Pressure (psi)

7500 BHIP - C-7.3.1.3

7000

6500

6000
0 50 100 150 200 250
Time (mins)

Figure 7.4 Comparison of simulation and field BHIP of Case 7.3.1.1-3- well A

8500
Field - BHIP - Well B
BHIP - C-7.3.1.1
BHIP - C-7.3.1.2
BHIP - C-7.3.1.3
Bottomhole Injection Pressure (psi)

8000

7500

7000

6500
0 50 100 150 200 250 300
Time (mins)

Figure 7.5 Comparison of simulation and field BHIP of Case 7.3.1.1-3 - well B
137
7.3.2 Effects of Limiting Length of Fracture Propagation

It is evident from the results of previous cases that decreasing the reduction factor will

not help us in history matching injection pressure. A mechanism that would create larger

pressure increase with time is required. One method is to restrict or confine fracture

propagation in length (half length), which means controlling fracture propagation by

modifying transmissibility of grid blocks in fracture plane only within an assumed

fracture half length. In this way fracture half length cannot exceed a specified number of

blocks for which transmissibilities are modified. Case 7.3.1.2 is selected as base case for

simulation runs in this section. Several simulation cases are run both for well A and B

using different values of pre determined fracture half length. Reservoir parameters used

in simulation run are same as base case. Permeability multipliers given in Figure 7.2 are

used everywhere except in fractured blocks for Case 7.3.2.1 and 7.3.2.2. Permeability

multipliers defined by stress factor of 3.0 are also used in fracture plane grid blocks only

for Case 7.3.2.2 – well B and are applied in x-direction while value of 1.0 is used for

multipliers in other two directions.

The parameters used for history matching of injection pressure for wells A and B are

summarized in Table 7.4. Comparison of simulation results of current cases and

bottomhole field injection pressure are presented in Figure 7.6 and 7.7 for well A and B

respectively followed by the discussion of the results.

138
Table 7.4 Parameters varied in injection Cases 7.3.2.1 – 2 – Well A and B

Well A Well B
Property
(Case 7.3.2.1) (Case 7.3.2.1)
Permeability reduction factor 0.01
Permeability multipliers in x, y and z-direction S = 6.0
Fracture half length, ft 75 150
Well B
(Case 7.3.2.2)
Permeability reduction factor 0.01
Permeability multipliers in x - direction S = 3.0
Permeability multipliers in y and z - direction S = 6.0
Fracture half length, ft 150

8000
BHIP - Field - Well A
BHIP - C-7.3.1.2
BHIP - C-7.3.2.1
7500
Bottomhole Injection Pressure (psi)

7000

6500

6000
0 50 100 150 200 250
Time (mins)

Figure 7.6 Comparison of simulation and field BHIP of Case 7.3.2.1- well A

In Case 7.3.1.2 of Figure 7.6, which is constructed for well A, fracture propagation was

allowed in y-direction without imposing any constraint on it but in Case 7.3.2.1 length of

139
fracture propagation was confined to its half length of 75 ft. Pressure rises approximately

by 300 psi after 70 minutes of injection in Case 7.3.2.1, which means that it takes 70

minutes for fracture to propagate to 75 ft (half length). Field pressure afterward keeps

increasing but change in computed pressure is very small. This is the best history match

that can be achieved with uncoupled injection simulation for well A.

10500
BHIP - Field - Well B
BHIP - C-7.3.1.2
BHIP - C-7.3.2.1
BHIP - C-7.3.2.2
9500
Bottomhole Pressure (psi)

8500

7500

6500
0 50 100 150 200 250 300
Time (mins)

Figure 7.7 Comparison of simulation and field BHIP of Case 7.3.2.1-2- well B

Results of Cases 7.3.1.2, 7.3.2.1 and 7.3.2.2 for well B are presented in Figure 7.7. In

Case 7.3.1.2 fracture propagation is allowed in y-direction as explained earlier but in

Case 7.3.2.1 fracture propagation is restricted to half length of 150 ft. The huge increase

in pressure to value of 10,000 psi in Case 7.3.2.1 is due to pressure build up as soon as

fracture propagation reaches its confinement limit. Small matrix permeability

perpendicular to fracture face also restricts fracture leak-off in x-direction and fracturing

140
pressure increases its value to its limit set to 10,000 psia in the simulator. To decrease

pressure build up during injection period and avoid operating at maximum constraint, one

can increase the permeability of matrix in x-direction so that leak-off can increase in x-

direction and delay fracture propagation in y-direction. For this reason Case 7.3.2.2 was

run for well B, in which permeability multipliers are applied in x-direction for fractured

blocks with S value of 3.0. Multipliers were used only in x-direction without effecting

permeability in other directions. This method somewhat improves the results, which can

be seen in Figure 7.7 – Case 7.3.2.2, but simulation still did not match field pressure

curves.

At this point Case 7.3.2.1 for well A and Case 7.3.2.2 for well B represent best history

matched cases for uncoupled injection scenarios. This depicts the limitation of uncoupled

model to history match injection pressure.

7.4 HISTORY MATCHING OF FIELD INJECTION PRESSURE USING A

FULLY COUPLED GEOMECHANICAL INJECTION MODEL

This section of chapter deals with history matching field injection pressure using a fully

coupled injection geomechanical model. In coupled geomechanical simulation, because

stresses are continuously computed, properties are updated at each time step in reservoir

simulator by taking as an input effective stress from geomechanical part of the simulator.

Therefore there is no need to modify stress data to correct for poroelastic effects. To run a

fully coupled geomechanical simulation the original in-situ stress data given in Table 7.1

is used to calculate transmissibility and permeability multipliers, which is function of

141
effective stresses. Transmissibility of grid blocks in fracture plane is a function of

minimum horizontal effective stress given by Equation (7.6) while matrix permeability

multiplier is function of mean effective stress in coupled geomechanical simulation

defined by Equation (6.3).

Where = Minimum horizontal effective stress

= Minimum horizontal total stress

= Biot’s constant

Minimum horizontal effective stress in Equation (7.6) is function of both time and space;

at any particular point in reservoir block it is constantly changing with time during

pressurization. Biot’s constant of 1.0 was initially used for effective stress calculation.

Run time for coupled simulation is very large and consequently a detailed study for each

parameter was not possible due to time constraints. The sensitivity study and calculations

shown here are performed for well A. Only conclusions and end results are then applied

to well B to get a history match. Because the results and conclusions obtained for well A

hold also true for well B, detailed discussion is not presented here. Permeability

multipliers calculated using original in-situ stresses are given in Figure 7.8 while

transmissibility multipliers for different fracture permeability reduction factors R fa are

presented in Figure 7.9 for well A. Same type of curves can be generated for well B using

the same equation and data set given in Table 7.1.

142
25,000
1,000
Injection - Well A - S=6

20,000 800

K(p)/Ki 600
15,000
K(p)/Ki

400

10,000
200

5,000 0
0 500 1000 1500 2000 2500 3000 3500 4000
Mean Effective Stress (psi)

0
0 1000 2000 3000 4000
Mean Effective Stress (psi)

Figure 7.8 Effective stress dependent permeability curve - coupled injection - well A

1.00E+08

1.00E+06

Transmissibility Multipliers
TmultY - Rfa=0.00001
TmultZ - Rfa=0.00001 1.00E+04
TmultY - Rfa=0.000007
TmultZ - Rfa=0.000007
TmultY - Rfa = 0.0000572
TmultZ - Rfa = 0.0000572 1.00E+02

1.00E+00
-3500 -3000 -2500 -2000 -1500 -1000 -500 0
Minimum Horizontal Effective Stress (psi)

Figure 7.9 Transmissibility multipliers for coupled injection cases - well A

143
Sensitivity study was performed on two history matching parameters, i.e., maximum

length of fracture propagation and Biot’s constant. Permeability multipliers and

transmissibility for fractured blocks were as presented in Figures 7.8 and 7.9.

7.4.1 Effects of Limiting Length of Fracture Propagation

Few simulation cases are run using different values of permeability reduction factor and

confining length of fracture propagation. For this purpose, it was assumed that rock

behaves as a perfectly elastic material which does not exhibit hysteresis during loading

and unloading. A base case here (Case 7.4.1.1) was the run allowing unlimited fracture

propagation in y-direction and modifying Biot’s constant value in the geomechanical

simulator to 0.65. Summary of history matching parameters varied in cases run in this

section are presented in Table 7.5.

Table 7.5 Parameters varied in coupled injection Cases 7.4.1.1 – 4 – Well A

Property – Well A Case 7.4.1.1 Case 7.4.1.2 Case 7.4.1.3 Case 7.4.1.4

Biot’s constant 0.65


Permeability reduction
0.00001 0.000007 0.00000572
factor Rfa
Fracture half length, ft Not restricted 100 150 135

In all the cases same Biot’s constant is used. The reason for using such a small value is

that, as will be shown in the next section, the use of high values of  (close to 1) creates

large poroelastic stresses, which prevent fracture from initiating in the model. The

objective here is to restrict or confine fracture propagation to an assumed fracture half

144
length by restricting fracture transmissibility modifiers to this length in the simulator. In

all cases the fracture height was assumed to be equal to pay zone thickness. Both

permeability multipliers and transmissibility are function of effective stress in coupled

modeling and entered in form of tables. All the numbers in Table 7.5 are used as an initial

guess for initial simulation for sensitivity study, which serves as base cases for further

simulation runs. Simulation results for all these cases and field injection pressure are

presented in Figure 7.10 for well A.

8500
BHIP - Field - Well A
BHIP - C-7.4.1.1
BHIP - C-7.4.1.2
BHIP - C-7.4.1.3
Bottomhole Injection Pressure (psi)

BHIP - C-7.4.1.4
7500

6500

5500
0 50 100 150 200 250
Time (mins)

Figure 7.10 Comparison of simulation and field BHIP of Cases 7.4.1.1-4 - well A

Fracture propagation was allowed in y-direction in base case (7.4.1.1) and there is a slow

continuous pressure build observed for that case due to uninterrupted fracture

propagation. The fracturing pressure rise is due to the pressure drop down in the fracture.

In case 7.4.1.2 the same parameters were used except fracture propagation was restricted
145
to half length of 100 ft. Pressure started building up after 70 minutes of injection, which

was the time taken by fracture to reach half length of 100 ft shown in Figure 7.10. The

restriction at the tip of fracture preventing fracture growth is causing pressure build up

(increment) upon further injection. It was revealed while analyzing the results in detail

that the downward dip between 100 – 110 minutes is due to numerical errors. Since this

case did no matches field pressure therefore no further investigation was carried out. In

Case 7.4.1.3 both permeability reduction factor and fracture half length were changed.

The effect of permeability reduction is discussed in detail in uncoupled simulation

section; decreasing its value shifts pressure injection curve upward which can be

observed in Figure 7.10 by comparing Cases 7.4.1.2, 7.4.1.3 and7.4.1.4. Limiting fracture

half length was increased to 150 ft from 100 ft. Very little bump in injection pressure is

now observed to be on the low side which is an indication that if all other parameters are

kept constant then the correct fracture half length will fall in between 100 to 150 ft to get

a reasonable match.

Case 7.4.1.4 was run by reducing further the permeability reduction factor (Rfa) and using

fracture half length of 135 ft as shown in Figure 7.10. Although simulation does not

exactly match field injection pressure, it is the best history match so far. It is concluded

that injection history match requires some mechanism to constrain fracture propagation at

a late stage. This issue was not pursued further; however, the coupled cases show much

improvement and are much better than the uncoupled simulations.

Discussion on the late time mismatch:

It is important to point out that history matching of field injection pressure after 190

minutes of injection cannot be achieved through our simulation results. It was observed

146
by thorough investigation of field treatment report that the injected proppant

concentration was increased after 190 minutes to approximately three times of the overall

average concentration. The snap shot of field treatment report is presented in Figure 7.11;

increase in BHIP after 190 minutes is due to increase in proppant concentration at that

time. As our simulation study does not include coupling of fracture propagation

simulation with proppant transport, modeling of fracture propagation based on downhole

variable proppant concentration is not possible here and beyond the scope of this study.

Fracture modeling in our case was performed based on total amount of downhole injected

slurry into formation. In light of above discussionTreatment Data Charts


it will be fair assumption to compare

simulation results COMPANY


BP CANADA ENERGY with field data for first 190 minutes of injection only. Gelled Water Frac
CO2 Crosslinked
BP Hz Noel a-A88-J 93-P-1 Cadomin (Gas) June 8, 2006

Wellhead Treating Bottomhole


70000

60000
Pressure ( kPa )

50000

40000

30000

20000

10000

0
0 50 100 150 200 250

Wellhead Inline Bottomhole


500
450
Concentration ( kg/m³ )

400
350
300
250
200
150
100
50
0
0 50 100 150 200 250

Blender Clean Blender Slurry Combined


14.0
Figure 7.11 Proppant concentration, BHIP and rate - well A
12.0
Rate ( m³/min )

Figure10.0
7.12 represents simulation and field bottomhole injection pressure for first 190
8.0

6.0
minutes of injection.
4.0

2.0
147
0.0
0 50 100 150 200 250
Elapsed Time ( min )
7500
BHIP - Field - Well A

BHIP - C-7.4.1.4
Bottomhole Injection Pressure (psi)

7000

6500

6000
0 50 100 150 200
Time (mins)

Figure 7.12 History matched case – Coupled simulation - well A

7.4.2 Effects of Biot’s Constant ( )

Sensitivity study was performed to see the effects of Biot’s constant on injection pressure

and here we only report the results for well A. Summary of the values of Biot’s constant,

used in two sub-cases run in this section is given in Table 7.6. History matched case from

previous section, i.e., Case 7.4.1.4 is used as a base case for these runs. All other

parameters are kept constant. Comparison of simulation results for the sub cases and the

base case together with the field injection pressure is presented in Figure 7.13.

148
Table 7.6 Parameters varied in injection cases 7.4.2.1 – 2 – Well A

Property – Well A Case 7.4.2.1 Case 7.4.2.2

Biot’s constant 0.75 1.0


Permeability reduction factor 0.00000572
Fracture half length, ft 135

11000

10000
Bottomhole Injection Pressure (psi)

BHIP - Field - Well A


BHIP - C-7.4.1.4
BHIP - C.7.4.2.1
9000
BHIP - C-7.4.2.2

8000

7000

6000
0 50 100 150 200 250
Time (mins)

Figure 7.13 Comparison of simulation and field BHIP of Cases 7.4.1-4 - well A

It is evident from Figure 7.13 that increasing Biot’s constant changes the results

tremendously. In the case of = 1 (Case 7.4.2.2) the model failed to initiate fracture

because the poroelastic stress component caused by injection pressure was too high and

the total stress increased above the injection pressure limit (set at 10,000 psia). The result

is a rapid, almost instant increase of pressure up to the injection pressure limit, which

149
plots on the graph as a vertical line close to time=0 followed by a horizontal line. The

smaller the Biot constant the slower is the increase in total stress and it is less difficult to

fracture the rock. Note that for fracture initiation (and propagation) minimum effective

stress must be negative; in other words injection pressure should be higher than minimum

horizontal total stress. Bottomhole injection pressure the wellbore grid block for fracture

# 2 (2nd fracture from the line of symmetry) is plotted vs. minimum horizontal effective

stress in Figure 7.14. It is evident from the curves that it is easier to create or propagate

fracture for smaller values of Biot’s constant. No fractures are created in Case 7.4.2.2

because the minimum horizontal effective stress is positive for all pressure values and

continuous injection with a BHIP limit will force a reduction in the injection rate in the

model. That is why injection pressure jumps to maximum value of 10,000 psi in very

early time and remains constant. It is therefore concluded that Biot’s constant equal to 1.0

is not reasonable assumption and should be less than 1.0. Biot’s constant can be

calculated from Equation (6.10) if rock bulk and grain modulus are known, which can be

measured in the laboratory. In absence of laboratory data approximate value can be

estimated by history matching which in our case led to the value of 0.65, but such

estimates must be used with necessary caution. Case 7.4.1.4 represents best history

matched case in coupled geomechanical simulation cases.

150
10000
Biot's Constant=0.65 - C-7.4.1.4
Biot's Constant=0.75 - C-7.4.2.1
Biot's Constant=1.0 - C-7.4.2.2
Bottomhole Injection Pressure (psi)

8000

6000

4000

2000
-3000 -2000 -1000 0 1000 2000 3000
Minimum Horizontal Effective Stress (psi)

Figure 7.14 Minimum effective stress and BHIP at different Biot’s constant

numbers – Coupled cases - well A

7.5 FAILURE PREDICTIONS

The simulations of the injection process presented in this Chapter showed that one must

assume a substantial stress-dependent enhancement of permeability around the primary

single plane fracture (SPF) to history match the injection pressures. Subsequently, the

permeability of the SRV will decline during production as was described in detail in

Chapter 6. Often it is postulated that the creation of this SRV is due to shear fracturing,

i.e., creating shear failure. Coupled modeling provides us with the tool to investigate

under what conditions shear fracturing occurs and what would be the extent of the SRV if

it was caused purely by shear failure. This aspect is examined in the present Section.

151
7.5.1 Tensile Failure

When tensile stress across a plane exceeds critical limit then tensile failure occurs. This

critical limit is called the tensile strength or ultimate tensile strength (UTS). It is a

characteristic property of rock and does not depend on size of specimen. It is opposite of

compressive strength and has units of stress. For linearly elastic material it can be

determined from stress strain curve in laboratory testing. Here all the compressive

stresses are denoted by positive sign. Tensile failure will occur when minimum effective

stress exceeds tensile strength of rock, is given as:

Where = Minimum effective stress

= Ultimate tensile strength or tensile strength

The tensile failure criterion is applied to determine the propagation of the main fracture

(SPF) through grid blocks. In some rare instances, tensile failure can also occur in the

reservoir around the SPF (e.g. due to thermal effects, Tran et al., 2012).

7.5.2 Shear Failure

When shear stresses along a plane in a specimen exceed shear strength of material then

shear failure occurs. The two sides of failure surface move relative to each other and

friction forces, which depend on normal forces acting on specimen, oppose this relative

movement. So shear strength of material / rock indirectly depends on the normal stress

acting over the failure plane.

152
There are different shear criterions available in literature such as Tresca, Mohr-Coulomb

and Griffith. For this study Mohr-Coulomb criterion is selected to predict failure

mechanism during injection. The objective of this study is to determine which type of

failure occurred during injection or pressurization.

(a) (b)

Figure 7.15 Schematics of tensile and shear failure under normal loading

7.5.3 Mohr – Coulomb Shear Failure Criterion

Mohr–Coulomb theory is a mathematical model describing the response of brittle

materials to shear stress as well as normal stress. It is a general and frequently used

criterion which postulates a linear relationship between shear stress at failure and normal

effective stress acting on failure plane and is given as:

Where = Shear stress at failure

= Shear strength or Cohesion

= Friction angle

153
= Effective normal stress

When friction angle is zero ( ) then Equation (7.7) becomes Tresca Criterion, so

Tresca criterion is special case of Mohr-Coulomb. Mohr-Coulomb criterion in graphical

form is presented in Figure 7.16.

Figure 7.16 Mohr-Coulomb failure criterion for critical stress state – A graphical

representation (from Faejer et al., 2008)

The distance of point of intersection and origin is denoted by and is called attraction. It

is related to other Mohr-Coulomb parameters by following relation.

and in Figure 7.16 is maximum and minimum effective stress. Mohr-Coulomb

criteria states that specimen fails under this criterion when circle touches failure line,

which means that failure criteria for some plane(s) is fulfilled. The value of intermediate

effective stress has no influence on failure criteria as it lies between maximum and

minimum effective stress so pure shear failure is independent of intermediate stress.

154
Figure 7.16 represents a case in which failure occurred in specimen referring it to critical

stress state. The shear and corresponding effective normal stress from Figure 7.16 is

Where = Failure plane orientation angle

gives orientation of failure plane with respect to maximum effective stress and it is

related to friction angle by the following relation.

If cohesion is known then uniaxial compressive strength, , can be calculated by given

relation.

To investigate if tensile or shear failure will occur; time-history of pressure and stresses

was extracted for specific grid blocks from a coupled simulation run. A program was

written in MATLAB® to present failure prediction of these grid blocks in graphical form

based on Mohr-Coulomb failure criterion. For this purpose, the history matched case, i.e.,

Case 7.4.1.4 was used. All the fractures behave the same way and pressure propagation is

also almost the same for all fractures. Therefore only one fracture is selected for analysis

which is fracture # 4 (4th fracture from line of symmetry) and conclusions are drawn from

this analysis will apply to all sets of fractures. Four grid blocks were selected and marked

as shown in Figure 7.17, which represents layer view of fracture in x-y plane. The well is

completed in x-direction and block 1 represents the perforation location.

155
Block 2

Block 3
Block 1 (Well) Block 4
Y

X
(Well )
Figure 7.17 Fluid pressure after 236 days and selected grid block locations – Well A

Minimum effective stress with respect to time for all four blocks is shown in Figure 7.18.

All the stress values are taken from the geomechanical module output for history matched

case and selected grid blocks. Minimum effective stress in block 1 falls to negative value

in very early time, which is an indication that fracture is created within this block.

156
3000
Block - 1 Block - 2

Block - 3 Block - 4
2000
Minimum Effective Stress (psi)

1000

-1000

-2000

-3000
0 50 100 150 200 250
Time (mins)

Figure 7.18 Minimum effective stress block 1 – 4 – history matched case – Well A

10000

8000
Total Stress(s) & BHIP (psi)

6000

Sx - Block 1
Sy - Block 1
Sz - Block 1
4000 Pressure - Block 1
Sx - Block 2
Sy - Block 2
Sz - Block 2
Pressure - Block 2
2000
0 50 100 150 200 250
Time (mins)

Figure 7.19 Total in-situ stresses and BHIP for block 1 and 2 –– Well A
157
Stress in block 2 is constant for first 40 minutes of injection then started falling off until

fracture is created at 68 minutes after injection (the point when minimum effective stress

becomes negative). Stresses both for block 3 and 4 are positive which indicates that no

tensile fractures are created in these blocks (as expected). Total in-situ stresses, i.e.,

stresses in x, y, and z-direction and bottomhole fluid injection pressure for block 1 and 2

are presented in Figure 7.19. Pressure and minimum total stress, i.e., Sx in block 2, is

constant until 47 minutes of injection and difference between them is positive, which is

an indication that no fractures are initiated till this time. But difference becomes negative

after 67 minutes of injection when pressure becomes larger than minimum horizontal

total stress, and fracture initiation or opening starts after this point.

Mohr-Coulomb failure criterion is used to construct the failure envelope and predict

failure based on minimum and maximum effective stress values from the coupled

simulation output and rock geomechanical data given in Table 3.3. As discussed earlier

fracture initiation or opening was observed only in block 1 and 2 for the whole injection

period. So failure envelope was constructed for only those blocks. Stresses values and

other properties used to construct Mohr-Coulomb are given in Table 7.7. Circle 1, 6 in

Table 7.7 represents 1st value of stresses for block 1 and 2 respectively. Circle 1 is for

block 1 and circle 6 is for block 2. Mohr – Coulomb failure envelope for stress values in

Table 7.7 is presented in Figure 7.20 for well A. Similar envelope can be constructed

from coupled simulations output for well B.

158
Table 7.7 Mohr-Coulomb circles input parameters – Block 1 and 2 – Well A

= 30°, = 46570 psi , = 13443 psi


Block 1 Block 2

Time (mins)
(psi) (psi) (psi) (psi)
11.51 2670 - 1461 4569 2504 Circle – 1, 6
70.30 2600 -1626 3168 -218.50 Circle – 2, 7
155.10 2265 -2329 2349 -2329 Circle – 3, 8
162.62 2439 -1924 2590 -1786 Circle – 4, 9
237.39 2237 -2296 2328 -2303 Circle – 5, 10

Mohr - Coulomb Failure Envelope


18000
Circle 1-1
Circle 2-1
16000 Circle 3-1
Circle 4-1
Circle 5-1
14000 Circle 6-2
Circle 7-2
Circle 8-2
12000
Circle 9-2
Circle 10-2
10000
Shear Stress

8000

6000

4000

2000

-2000
-2.5 -2 -1.5 -1 -0.5 0 0.5
Effective Normal Stress 4
x 10

Figure 7.20 Mohr – Coulomb failure envelope for Co= 46570psi – Well A

It is evident from Figure 7.20 that there is no shear failure during injection because

Mohr’s circle is way below the failure line. It is obvious that the dominant failure

159
mechanism in these blocks is tensile because the SPF penetrated them. Although the data

shown covers only some selected times, we also repeated the same exercise for all time

steps and confirmed that no shear failure occurred. Complete spatial map of the failure

can be obtained by plotting “stress level” S L, a feature offered in GeoSim®

(Geomechanical Simulator) which represents the ratio of the size of the Mohr circle at

any point to the circle at critical state, i.e., when the circle touches the failure line. Stress

level therefore ranges between 0 and 1. When SL < 1 there is no shear failure, and 1

represent when circle touches failure line, shear failure. Stress level for fracture # 4 after

237 minutes of injection (end of injection) is shown in Figure 7.21 for YZ crossection

through the fracture plane followed by XY areal view for the same fracture block and

injection time in Figure 7.22. There is no shear failure anywhere.

Block 2 Well

Figure 7.21 Stress level in Fracture # 4 after 237mins of injection – Co= 46570 psi –

YZ cross section – Well A

160
Well

Figure 7.22 Stress level in Fracture # 4 after 237mins of injection – Co= 46570 psi –

XY areal view – Well A

Two more cases were run by reducing the uniaxial compressive strength by 5 and 10

times to a base value given in Table 7.7 while keeping other parameters such friction

angle, elastic modulus and Poisson’s ratio the same as in the previous case. Mohr-

Coulomb circles for these cases are shown in Figure 7.23 and 7.26 while stress level is

shown in Figure 7.24, 7.25, 7.27 and 7.28. Circle – 10 and circle – 8 almost touch the

failure line in Figure 7.23, and stress level values both for block – 1and 2 are equal to

0.98 - very close to failure. By reducing the uniaxial compressive strength further by

factor of 2, majority of circles in Figure 7.26 touches the failure line, indicating shear

failure. The results can be confirmed both in Figure 7.27 and 7.28 as stress level are

much larger than 1, it is also important to point out that simulation for last case was run

for 141 mins as compared to other cases, which were run for 237 mins.

161
Mohr - Coulomb Failure Envelope
6000
Circle 1-1
Circle 2-1
Circle 3-1
5000 Circle 4-1
Circle 5-1
Circle 6-2
Circle 7-2
4000 Circle 8-2
Circle 9-2
Circle 10-2
Shear Stress

3000

2000

1000

-1000
-5000 -4000 -3000 -2000 -1000 0 1000 2000 3000 4000 5000
Effective Normal Stress

Figure 7.23 Mohr – Coulomb failure envelope for Co = 9314 psi – Well A

Block 2 Well

Figure 7.24 Stress level in Fracture # 4 after 237 mins of injection – Co= 9314 psi –
YZ cross section – Well A
162
Well

Figure 7.25 Stress level in Fracture # 4 after 237 mins of injection – Co= 9314 psi –
XY areal view – Well A

Mohr - Coulomb Failure Envelope


4000
Circle 1-1
Circle 2-1
3500 Circle 3-1
Circle 4-1
Circle 5-1
Circle 6-2
3000
Circle 7-2
Circle 8-2
Circle 9-2
2500
Circle 10-2
Shear Stress

2000

1500

1000

500

-500
-3000 -2000 -1000 0 1000 2000 3000 4000 5000
Effective Normal Stress

Figure 7.26 Mohr – Coulomb failure envelope for Co= 4657psi –– Well A
163
Block 2 Well

Figure 7.27 Stress level in Fracture # 4 after 141 mins of injection – Co= 4657 psi –
YZ cross section – Well A

Well

Figure 7.28 Stress level in Fracture # 4 after 141 mins of injection – Co= 4657 psi –
XY areal view – Well A

164
From above simulation runs, Mohr-Coulomb failure analysis and knowledge of presence

of micro cracks and heterogeneities in tight sands as recorded by microseismic, it is

concluded that the high original value of uniaxial compressive strength (which does not

allow any shear events) is unlikely. Reducing the C0 to account for weak planes and

natural fractures then will predict possibility of shear fracturing and shear-generated SRV

creation.

7.6 SUMMARY

 The fracture permeability factor (Rfa) in injection model has little effect on the

slope of pressure. However, it shifts the pressure curve upward.

 Confining the fracture propagation helped to increase the pressure and thus nearly

matches the field data.

 Decreasing value of Biot’s constant results in a decrease of effective stresses. For

larger values of Biot’s constant it is very difficult to fracture the formation.

 Correct value of Biot’s constant should be used as it affects the results, especially

in matching injection pressure. This should be done using laboratory

measurement of bulk and grain modulus or triaxial testing.

 No shear events were predicted in injection cases if a high value of uniaxial

compressive strength of rock is used. However, shear failure is detected if a lower

(effective) value of UCS is used.

 Microseismic events detected in the field during stimulation can be attributed to

opening of existing micro cracks in reservoir or creating shear failure.

165
CHAPTER EIGHT: RESULTS AND CONCLUSIONS

The following main conclusions are drawn from the studies presented in previous

chapters:

For production history matching:

 The method for calculating stress changes presented in Chapter 6 proves to be a

good approximation to capture poroelastic effects and can be used in conventional

reservoir simulation without running fully coupled geomechanical simulation.

 Moving from uncoupled to fully coupled geomechanical modeling without

correcting data sets for poroelastic effects will results in higher production. Data

sets must be corrected for stress changes in the reservoirs. Constant mean total

stress assumption is not a realistic approach and will always results in

overestimated production. Under depletion, the total mean stress is decreasing due

to poroelasticity.

 Using baseline propped fracture conductivity in simulation and other calculations

will always over estimate results; so realistic proppant conductivity should be

used.

 Net pay thickness has little effects on production in early time but affects

production at late time.

 Size and permeability of Stimulated Reservoir Volume (SRV) has significant

effects on results. Permeability inside the SRV must be larger than formation

permeability, and both are stress- or pressure-dependent during production.

166
 An accurate history match is impossible without use of pressure / stress dependent

permeability.

 Length of Stimulated Reservoir Volume (SRV) is generally much larger than

propped fracture length.

 The dimensions of SRV are different for x-linked and slick water fracturing wells

– shorter for x-linked fracture job.

 A realistic and accurate way to obtain a history match is the use of pressure

dependent permeability multipliers both inside and outside SRV.

 Permeability of the region inside SRV has larger pressure dependence than region

outside SRV due to presence of complex network of induced fractures.

 The two wells seem to have somewhat different virgin reservoir permeability.

For modeling of the fracturing process:

 Permeability reduction factor (Rfa) in injection model has no effects on results. It

shifts the pressure curve upward

 Confining the length of fracture propagation causes to increase of pressure in later

part of the job and thus improves the matches to field pressure.

 Decreasing the value of Biot’s constant results in decrease of effective stresses.

For larger values of Biot’s constant it is very difficult to fracture the formation.

 No shear events are detected in injection cases when a high value of uniaxial

compressive strength of the rock (UCS) is assumed, representative of intact rock.

167
 Shear failure regions around the fractures are predicted when the UCS is lowered

to represent media with pre-existing fractures or planes of weakness.

 Microseismic events detected on field can be attributed to opening of existing

micro cracks in reservoir or secondary shear fracturing. Therefore the increase of

permeability in the SRV can be a combination of shear fractures and stress

sensitivity of the matrix.

Our conclusions are broadly in agreement with laboratory experiments and petro physical

data on tight gas sands, and with the interpretation of microseismic data. They are also

supported by theoretical correlations for stress-sensitive permeability and results of other

geomechanical modeling.

168
REFERENCES

American Petroleum Institute: “Recommended Practices for Evaluating Short Term Proppant
Pack Conductivity”, API RP 61, Oct. 1989.

Anderson, R. A., Ingram, D. S., and Zanier, A. M. 1973. Determining Fracture Pressure
Gradients from Well Logs. J. Pet Tech 25 (11): 1259-1268. SPE-4135-PA.

Barree, R. D., Cox, S. A., Barree, V. L. et al. 2003. Realistic Assessment of Proppant Pack
Conductivity for Materials Selection. Paper SPE 84306 presented at the SPE Annual
Technical Conference and Exhibition, Denver, Colorado, 5-8 October.

Bell, J. S., and Babcock, E. A. 1986. The Stress Regime of the Western Canadian Basin and
Implications for Hydrocarbons Production. Bull. Of Cdn. Pet. Geo. 34 (3): 364-378.

Bell, J. S., Price, P. R., and McLellan, P. J. 1994. In-situ stress in the western Canada
Sedimentary Basin. In Geological atlas of the western Canada sedimentary Basin.
Mossop, G.D., and Shetsen, I.: Canadian Society of Petroleum Geologists and Alberta
Research Council, 439 – 446.

Biot, M.A. 1941. General Theory of Three Dimensional Consolidation. J. Applied Physics 12:
155-164.

Breckels, I. M., and Van Eekelen, H. A. M. 1982. Relationship between Horizontal Stress and
Depth in Sedimentary Basins. J. Pet Tech 34 (9): 2191-2199. SPE-10336-PA.

Brighenti, G. 1989. Effect of Confining Pressure on Gas Permeability of Tight Sandstones.


Paper IS-1989-024 presented at the ISRM International Symposium, Pau, France, Aug.
30 – Sept. 2.

Brower, K. R., and Morrow, N. R. 1985. Fluid Flow in Cracks as Related to Low-Permeability
Gas Sands. SPE Journal 25 (2): 191-201. SPE-11623-PA.

169
Buchsteiner, H., Warpinski, N. R., and Economides, M. J.1993. Stress Induced Permeability
Reduction in Fissured Reservoirs. Paper SPE 26513 presented at the SPE Annual
Technical Conference and Exhibition, Houston, Texas, 3 – 6 October.

Carpenter, C. B., and Spencer, G. B. 1940. Measurements of Compressibility of Consolidated


Oil Bearing Sandstones. U.S. Bureau of Mines, R. I. 3450.

Carr, N. L., Kobayashi, R., and Burrows, D. B. 1954. Viscosity of Hydrocarbon Gases under
Pressure. J. Pet Tech 6 (10): 47-55. SPE-297-G.

Cater, R. D. Derivation of the General Equation for Estimating the Extent of the Fractured
Area. Appendix to: Howard, G. C., and Fast, C. R. 1957. Optimum Fluid
Characteristics for Fracture Extension. Drilling and Production Practice, API 261-270.

Cinco-Ley, H., and Samaniego-V, F. 1981. Transient Pressure Analysis for Fractured Wells. J.
Pet Tech 33 (9): 1749-1766. SPE-7490.

Clark, J. B. 1949. A Hydraulic Process Increasing the Productivity of Wells. Trans. AIME,
186, 1-8.

Cleary, M. P., Keck, R. G., and Mear, M. E. 1983. Microcomputer Models for the Design of
Hydraulic Fractures. Paper SPE 11628 presented at the SPE/DOE Low Permeability
Gas Reservoirs Symposium, Denver, Colorado, 14-16 March.

Cleary, M. P., and Lam, K. Y. 1983. Development of Fully a Three-Dimensional Simulator for
Analysis and Design of Hydraulic Fracturing. Paper SPE 11631 presented at the
SPE/DOE Low Permeability Gas Reservoirs Symposium, Denver, Colorado, 14-16
March.

Clifton, R. J., and Abou-Sayed, A. S. 1981. A Variational Approach to the Prediction of the
Three Dimensional Geometry of the Hydraulic Fractures. Paper SPE 9879 presented at
the SPE/DOE Low Permeability Gas Reservoirs Symposium, Denver, Colorado, 27-29
May.

170
Davies, J. P., and Davies, D. K. 2001. Stress Dependent Permeability: Characterization and
Modeling. SPE Journal 6 (2): 224-235. SPE-71750-PA.

Dedurin, A. V., Majar, V. A., Voronkov, A. A. et al. 2006. Designing Hydraulic Fractures in
Russian Oil and Gas Fields to Accommodate Non-Darcy and Multiphase Flow. Paper
SPE 101821 presented at the SPE Russian Oil and Gas Technical Conference and
Exhibition, Moscow, 3-6 October.

Dobrynin, V. M. 1962. Effect of Overburden Pressure on Some Properties of Sandstones. SPE


Journal 2 (4): 360-366. SPE-461-PA.

Dranchuck, P. M., and Abu-Kassem, J. H. 1975. Calculation of Z Factors for Natural Gases
Using Equations of State. J. Cdn. Pet. Tech. 14 (3): 34-36. SPE-75-03-03.

Dranchuck, P. M., Purvis, R. A., and Robinson, D. B., 1974. Computer Calculation of Natural
Gas Compressibility Factors Using the Standing and Katz Correlation. Paper SPE 73-
112 presented at the SPE Annual Technical Meeting, Edmonton, 8-12 May.

Economides, M. J., and Nolte, K .G. 2000. Reservoir Simulation, Third Ed. John Wiley &
Sons, Inc.

Fatt, I. 1953. The Effect of Overburden Pressure on Relative Permeability. Trans. AIME, 198,
325-326.

Fatt, I., and Davis, D. H. 1952. Reduction in Permeability with Overburden Pressure. Trans.
AIME, 195, 329.

Fjær, E., Holt, R. M., Horsrud, P. et al. 2008. Petroleum Related Rock Mechanics, 2nd Edition.
Elsevier. 21-36.

Gangi, A. F. 1978. Variations of Whole and Fractured Porous Rock Permeability with
Confining Pressure. Int. J. Rock. Mech. Min. Sci. & Geomech. Abstr. 15: 249-257.

171
Geertsma, J., and de Klerk, F. 1969. A Rapid Method of Predicting Width and Extent of
Hydraulically Induced Fractures. J. Pet Tech 21 (12): 1571-1581. SPE-2458-PA.

Gray, D. H., Fatt, I., and Bergamini, G. 1963. The Effect of Stress on Permeability of
Sandstone Cores. SPE Journal 3 (2): 95-100. SPE-531-PA.

Hall, H. N. 1953. Compressibility of Reservoir Rocks. Trans. AIME, 198, 309-311.

Harrison, E., Kieschnick, W. F., and McGuire, W. J. 1953. The Mechanics of Fracture
Induction and Extension. Trans. AIME, 201, 252-321.

Holditch, S. A. 2006. Tight Gas Sands. Distinguished Author Series, J. Pet Tech 58 (6): 86-93.
SPE-103356.

Hubbert, M. K., and Willis, D. G. 1957. Mechanics of Hydraulic Fracturing. Trans. AIME,
210, 153-168.

International Organization for Standardization: “Procedures for Measuring the Long Term
Conductivity of Proppants”, ISO 13503-5, July, 2006.

Jacob, C. E. 1940. On the Flow of Water in an Elastic Artesian Aquifers. Trans. AGU, 574-
586.

Jacob, C. E. 1950. Engineering Hydraulics, New York: John Wiley & Sons, Inc. 321-386.

Jennings, J. B., Carroll, H. B., and Raible, C. J. 1981. The Relationship of Permeability to
Confining Pressure in Low Permeability Rock. Paper SPE 9870 presented at the
SPE/DOE Low Permeability Gas Reservoirs Symposium, Denver, Colorado, 27-29
May.

Ji, L., Settari, A., Orr, D. W. et al. 2004a. Methods for Modelling Static Fractures in Reservoir
Simulation. Paper SPE 2004-260 presented at the SPE Canadian International
Petroleum Conference, Calgary, Alberta, 8-10 June.

172
Ji, L. Settari, A. Sullivan, R. B. et al. 2004b. Methods for Modeling Dynamic Fractures in
Coupled Reservoir and Geomechanics Simulation. Paper SPE 90874 presented at the
SPE Annual Technical Conference and Exhibition, Houston, Texas, 26 – 29 September.

Ji, L. Settari, A. Sullivan, and R. B. 2009. A Novel Hydraulic Fracturing Model Fully Coupled
with Geomechanics and Reservoir Simulation. SPE Journal 14 (3): 423-430. SPE-
110845-PA.

Jones, F. O. and Owens, W. W. 1980. A Laboratory Study of Low-Permeability Gas Sands. J.


Pet Tech 32 (9): 1631-1640. SPE-7551-PA.

Kawata, Y. and Fujita, K. 2001. Some Publications of Possible Unconventional Hydrocarbon


Availability Until 2100. Paper SPE 68755 presented at the SPE Asia Pacific Oil and
Gas Conference, Jakarta, 17-19 April.

Khristianovic, S. A., and Zheltov, Y. P. 1955. Formation of Vertical Fractures by Means of


Highly Viscous Liquid. Proc., Fourth World Pet. Cong., Rome, Sec. 2, 579-586.

Lee, A. L., Gonzalez, M. H., and Eakin, B. E. 1966. The Viscosity of Natural Gases. J. Pet
Tech 18 (8): 997-1000. SPE-1340-PA.

Lei, Q., Xiong, W., Yuan, J. et al. 2007. Analysis of Stress Sensitivity and Its Influence on Oil
Production from Tight Reservoirs. Paper SPE 111148 presented at the SPE Eastern
Regional Meeting, Lexington, Kentucky, 17 – 19 October.

Lin, J., and Zhu, D. 2012. Predicting Well Performance in Complex Fracture Systems by Slab
Source Method. Paper SPE 151960 presented at SPE Hydraulic Fracturing Technology
Conference, The Woodlands, Texas, 6-8 February.

Liu, X., and Civan, F. 1995. Formation Damage by Fines Migration Including Effects of Filter
Cake, Pore Compressibility, and Non-Darcy Flow - A Modeling Approach to Scaling
from Core to Field. Paper SPE 28980 presented at the SPE International Symposium on
Oilfield Chemistry, San Antonio, Texas, 14 – 17 February.

173
Liu, Y., and Sharma, M. M. 2005. Effect of Fracture Width and Fluid Rheology on Proppant
Settling and Retardation; An Experimental Study. Paper SPE 96208 presented at the
SPE Annual Technical Conference and Exhibition, Dallas, Texas, 9-12 October.

Liu, Y., Gadde, P. B., and Sharma, M. M. 2006. Proppant Placement using Reverse-Hybrid
Fracs. Paper SPE 99580 presented at the SPE Gas Technology Symposium, Calgary,
Alberta, 15 – 17 May.

Lorenz, J. C. 1999. Stress-Sensitive Reservoirs. J. Pet Tech: 61-63.

Masters, J.A. 1979. Deep Basin Gas Trap, Western Canada. AAPG Bulletin 63 (2): 152-181.

Mayerhofer, M. J., and Meehan, D. J. 1998. Waterfracs - Results from 50 Cotton Valley Wells.
Paper SPE 49104 presented at the SPE Annual Technical Conference and Exhibition,
New Orleans, Louisiana, 27 – 30 September.

McGuire, W. J. and Sikora, V. J. 1960. The Effect of Vertical Fractures on Well Productivity.
J. Pet Tech 12 (10): 72-74. SPE-1618-G.

McKee, C. R., Bumb, A. C., and Koenig, R. A. 1988. Stress-Dependent Permeability and
Porosity of Coal and Other Geologic Formations. SPE Form Eval 3 (1): 81-91. SPE-
12858-PA.

McLatchie, A. S., Hemstock, R. A., and Young, J. W. 1958. The Effect of Compressibility of
Reservoir Rock and Its Effects on Permeability. J. Pet Tech 10 (6): 49-51. SPE-894-G.

Meinzer, O. E. 1928. Compressibility and Elasticity of Artesian Aquifers. Econ. Geol. 23: 263-
271.

Montgomery, C. T. and Steanson, R. E. 1985. Proppant Selection: The Key to Successful


Fracture Stimulation. J. Pet Tech 37 (12): 2163-2172. SPE-12616.

Mossop, G. D. and Shetsen, I. (comp.). 1994. Geological atlas of the Western Canada
Sedimentary Basin; Canadian Society of Petroleum Geologists and Alberta Research

174
Council, URL <http://www.ags.gov.ab.ca/publications/wcsb_atlas/atlas.html>, [August
27, 2012].

Much, M. G. and Penny, G. S. 1987. Long Term Performance of Proppants under Simulated
Reservoir Conditions. Paper SPE 16415 presented at the SPE Low Permeability
Reservoirs Symposium, Denver, Colorado, 18-19 May.

Nassir, M., Settari, A., and Wan, R. G. 2012. Prediction and Optimization of Fracturing in
Tight Gas and Shale Using a Coupled Geomechanical Model of Combined Tensile and
Shear Fracturing. Paper SPE 152200 presented at the SPE Hydraulic Fracturing
Technology Conference, The Woodlands, Texas, 6-8 February.

Newberry, B. M., Nelson, R. F., and Ahmed, U. 1985. Prediction of Vertical Hydraulic
Fracture Migration Using Compressional and Shear Wave Slowness. Paper SPE 13895
presented at the SPE/DOE Low Permeability Gas Reservoirs Symposium, Denver,
Colorado, 19-22 March.

Nordgren, R. P. 1972. Propagation of a Vertical Hydraulic Fracture. SPE Journal 12 (4): 306-
314. SPE-3009-PA.

Olson, J. E., Bahorich, B., and Holder, J. 2012. Examining Hydraulic Fracture – Natural
Fracture Interaction in Hydrostone Block Experiments. Paper SPE 152618 presented at
SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 6-8
February.

Olson, J. E., and Wu, K. 2012. Sequential vs. Simultaneous Multizone Fracturing in Horizontal
Wells: Insights From a Non-Planar, Multifrac Numerical Model. Paper SPE 152602
presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands,
Texas, 6-8 February.

Ostensen, R. W. 1983. Microcrack Permeability in Tight Gas Sandstone. SPE Journal 23 (6):
919-927. SPE-10924-PA.

175
Ostensen, R. W. 1986. The Effect of Stress Dependent Permeability on Gas Production and
Well Testing. SPE Formation Evaluation 1 (3): 227-235. SPE-11220-PA.

Palisch, T., Duenckel, R., Bazan, L. et al. 2007. Determining Realistic Fracture Conductivity
and Understanding its Impact on Well Performance – Theory and Field Examples.
Paper SPE 106301 presented at the SPE Hydraulic Fracturing Technology Conference,
College Station, Texas, 29-31 January.

Palisch, T. T., Vincent, M. C., and Handren, P. J. 2010. Slickwater Fracturing: Food for
Thought. SPE Prod & Oper 25 (3): 327-344. SPE-115766-PA.

Pearson, C. M. 2001. Dimensionless Fracture Conductivity: Better Input Vales Make Better
Wells. J. Pet Tech 53 (1): 59-63. SPE-60184.

Perkins, T. K., and Kern, L. R. 1961. Widths of Hydraulic Fractures. J. Pet Tech 13 (9): 937-
949. SPE-89-PA.

Prats, M. 1961. Effect of Vertical Fractures on Reservoir Behavior – Incompressible Fluid


Case. SPE Journal 1 (2): 103-118. SPE-1575-G.

Reyes, L., and Osisanya, S. O. 2002. Empirical Correlation of Effective Stress Dependent
Shale Rock Properties. J. Cdn. Pet. Tech. 41 (12). SPE-02-12-02.

Rogner, H. H. 1996. An assessment of World Hydrocarbon Resources. IIASA, WP-96-26,


Luxemburg, Austria, May 1996.

Rushing, J. A., and Sullivan, R. B. 2003. Evaluation of a Hybrid-Frac Stimulation Technology


in the Bossier Tight Gas Sand Play. Paper SPE 84394 presented at the SPE Annual
Technical Conference and Exhibition, Denver, Colorado, 5-8 October.

Sampath, K., and Keighin, C. W. 1982. Factors Affecting Gas Slippage in Tight Sandstones of
Cretaceous Age in the Uinta Basin. J. Pet Tech 34 (11): 2715-2720. SPE-9872-PA.

176
Schubarth, S. K., Spivey, J. P., and Huckabee, P. T. 2006. Using Reservoir Modeling To
Evaluate Stimulation Effectiveness in Multilayered Tight Gas Reservoirs: A Case
History in the Pinedale Anticline Area. Paper SPE 100574 presented at the SPE Gas
Technology Symposium, Calgary, Alberta, 15 –175 May.

Settari, A., Puchyr, P. J., and Bachman, R. C. 1990. Partially Decoupled Modeling of
Hydraulic Fracturing Process. SPE Prod Eng 5 (1): 37-44. SPE-16031.

Settari, A. Sullivan, R. B., Walters, D. A. et al. 2002a. 3-D Analysis and Prediction of
Microseismicity in Fracturing by Coupled Geomechanical Modeling. Paper SPE 75714
presented at the SPE Gas Technology Symposium, Calgary, Alberta, 28 April – 2 May.

Settari, A. Sullivan, R. B., and Bachman, R. C. 2002b. The Modeling of the Effect of Water
Blockage and Geomechanics in Waterfrac. Paper SPE 77600 presented at the SPE
Annual Technical Conference and Exhibition, San Antonio, Texas, Sept. 29 – Oct 2.

Settari, A., Bachman, R. C., and Walters, D. A. 2005. How to Approximate Effects of
Geomechanics in Conventional Reservoir Simulation. Paper SPE 97155 presented at
the SPE Annual Technical Conference and Exhibition, Dallas, Texas, 9-12 October.

Settari, A., Sullivan, R. B., Rother, R. et al. 2009. Comprehensive Coupled Modeling Analysis
of Stimulations and Post-Frac Productivity – Case Study of the Wamsutter Field. Paper
SPE 119394 presented at the SPE Hydraulic Fracturing Technology Conference, The
Woodlands, Texas, 19–21 January.

Sharma, M. M., Gadde, P. B., Sullivan, R. B. et al. 2004. Slick Water and Hybrid Fracs in the
Bossier: Some Lessons Learnt. Paper SPE 89876 presented at the SPE Annual
Technical Conference and Exhibition, Houston, Texas, 26 – 29 September.

Sharma, M. M., Gadde, P. B., Liu, Y. et al. 2005. Selecting Strategies for Improving
Performance in Tight Gas Sands. Gas Tips 11 (4): 8-11.

177
Sneddon, I. N., and Lowengrud, M. 1969. Crack Problems in the Classical Theory of
Elasticity, 20-30. New York: SIAM Series in Applied Mathematics, John Wiley &
Sons.

Sutton, R. P. 1985. Compressibility Factors for High-Molecular-Weight Reservoir Gases.


Paper SPE 14265 presented at the SPE Annual Technical Conference and Exhibition,
Las Vegas, Nevada, 22 – 25 September.

Tao, Q., Ehlig-Economides, C. A., and Ghassemi, A. 2009. Investigation of Stress-Dependent


Fracture Permeability in Naturally Fractured Reservoir Using a Fully Coupled
Poroelastic Displacement Discontinuity Model. Paper SPE 124745 presented at the SPE
Annual Technical Conference and Exhibition, New Orleans, Louisiana, 4-7 October.

Thomas, R. D., and Ward, D. C. 1972. Effect of Overburden Pressure and Water Saturation on
Gas Permeability of Tight Sandstone Cores. J. Pet Tech 24 (2): 120-124. SPE-3634-
PA.

Urbancic, T., Baig, A., and Goldstein, S. 2012. Assessing Stimulation of Complex Natural
Fractures as Characterized Using Microseismicity: An Argument the Inclusion of Sub-
Horizontal Fractures in Reservoir Models. Paper SPE 152616 presented at SPE
Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 6-8 February.

Vairogs, J., Hearn, C. L., Dearing, D. W. et al. 1971. Effect of Rock Stress on Gas Production
from Low Permeability Reservoirs. J. Pet Tech 23 (9): 1161-1167. SPE-3001-PA.

Vairogs, J., and Rhoades, V. W. 1973. Pressure Transient Tests in Formations Having Stress
Sensitive Permeability. J. Pet Tech 25 (8): 965-970. SPE-4050-PA.

Vermylen, J. P., and Zoback, M. D. 2011. Hydraulic Fracturing, Microseismic Magnitudes,


and Stress Evolution in the Barnett Shale, Texas, USA. Paper SPE 140507 presented at
SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 24-26
January.

178
Walls, J. D. 1982. Tight Gas Sands-Permeability, Pore Structure, and Clay. J. Pet Tech 34 (11):
2708-2714. SPE-9871-PA.

Walsh, J. B. 1981. Effect of Pore Pressure and Confining Pressure on Fracture Permeability.
Int. J. Rock. Mech. Min. Sci. & Geomech. Abstr. 18: 429-435.

Warpinski, N. R., and Teufel, L. W. 1987. Influence of Geologic Discontinuities on Hydraulic


Fracture Propagation. J. Pet Tech 39 (2): 209-220. SPE-13224-PA, SPE – 17011, and
SPE – 17074.

Warpinski, N. R. 1991. Hydraulic Fracturing in Tight, Fissured Media. J. Pet Tech 43 (2): 146-
151, 208-209. SPE-20154-G.

Warpinski, N. R., Lorenz, J. C., Branagan, P. T. et al. 1993. Examination of a Cored Hydraulic
Fracture in a Deep Gas Well. SPE Prod & Fac 8 (3): 150-158. SPE-22876-PA, SPE –
26302, and SPE – 26946.

Weng, X., Kresse, O., Cohen., C. et al. 2011. Modeling of Hydraulic-Fracture-Network


Propagation in a Naturally Fractured Formation. SPE Prod & Oper 26 (4): 368-380.
SPE-140253-PA.

Wu, Y. S., Rutqvist, J., Karasaki, K. et al. 2008. A Mathematical Model for Rock
Deformation’s Effect on Flow in Porous and Fractured Reservoirs. Paper SPE 08-142
presented at the 42nd US and 2nd Canadian Rock Mechanics Symposium (UCRMS), San
Francisco, California, June 29 – July 2.

Wu, R., Kresse, O., Weng, X. et al. 2012. Modeling of Interaction of Hydraulic Fractures in
Complex Fracture Networks. Paper SPE 152052 presented at SPE Hydraulic Fracturing
Technology Conference, The Woodlands, Texas, 6-8 February.

Wyble, D. O. 1958. Effect of Applied Pressure on the Conductivity, Porosity and Permeability
of Sandstones. J. Pet Tech 10 (11): 57-59. SPE-1081-G.

179
Yarborough, L., and Hall, K. R. 1974. How to Solve Equation-of-State for Z-factors. Oil and
Gas Journal: 86-88.

180
APPENDIX A: SELECTED DATA FILES

GeoSim requires three input files to run a coupled problem. These files have different

extensions and are used for different parts of model.

Case – 6.6.dat : The Input File for Flow Model


'
' **** TERASIM RESERVOIR SIMULATOR ****
'
' ==== ELEMENT OF SYMMETRY MODEL ====
' Stimulated Well - no water (comparison run)
' Gas Reservoir with Inactive Bottom Water
' Cartesian COORDS
'
ATITLE= DRY GAS WELL A
' ====================================================================================
' MAIN CONTROL
' ====================================================================================
'
' KEYWORD VALUE
' ------- -------
GEOM= 3-D ' AREAL FOR SINGLE LAYER, 3-D FOR MULTI-LAYER
UNITS= FIELD ' FIELD OR SI
NINE= OFF ' LEAVE OFF FOR FRAC PROBLEMS
INIT= ON ' INITIALIZATION: INIT = ON OFF
NIN= 0 ' RESTART CONTROL: NIN = (LAST COMPLETED TS)
ALIGN= OFF ' USED ONLY FOR NESTED GRID PROBLEMS
IREG= OFF ' NO. OF EQUILIBRIUM REGIONS FOR INITIALIZATION
AREG= OFF ' NO. OF ACCOUNTING REGIONS
MROCK= OFF ' MULTIPLE ROCK TYPES
PFRAC= OFF ' PRESSURE DEPENDENT WATER RELATIVE PERMEABILITIES
HORIZ-VERT ' ENTER BOTH HORIZONTAL AND VERTICAL PERMS
GAS= DRY ' BLACK OIL MODE ON OR OFF
WATER= ON
OIL= NONE
CNARRAY
PXMULT = ON
PYMULT = ON
PZMULT = ON
END
' ====================================================================================
' LINK TO GEOSIM - GEOMECHANICAL SIMULATOR
' ====================================================================================
GMECH = EXPL
'
' DIMENSIONS
'
NWMAX=1 ' MAXIMUM NUMBER OF WELLS
NLMAX=8 ' MAXIMUM NUMBER OF COMPLETIONS (PERFORATIONS) PER WELL
END

181
' ====================================================================================
‘RESERVOIR DESCRIPTION - (GRID LISTS)
' ====================================================================================
' ==== NUMBER OF GRID CELLS: FUNDAMENTAL GRIDS ====
' X-DIREC Y-DIREC(V) Z-DIREC(W)
N1=110 N2=43 N3=17
'
' ==== X - DIREC GRID DIMENSIONS (N=1, NX) ====
'
+I=1 2 3 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5
+I=43 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10
+I=84 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 150 150 350 500 500 1000 1000 1000 1000 1000 1000 1000
END
'
' ==== y - DIREC GRID DIMENSIONS (N=2, NY) ====
'
+J=1 1000 1000 500 200 100 60 22 22 18 12 10 10 10 8 5 4 4 4 4 4 2 2 2 4 4 4 4 4 5 8 10 10 10 12 18 22 22 60 100
+J=40 200 500 1000 1000
END
'
' ==== Z - DIREC GRID DIMENSIONS (N=3, NZ) ====
'
+K=1 14 10 10 5 4 2 2 2 2 2 2 2 4 5 10 10 14
END
'
' ==== DEPTH TO THE TOP OF RESERVOIR (ONE FOR EACH GRID FOR TOP LAYER) ====
'
I=1/* J=1/* K=1/1 5250.0
END
'
' ==== POROSITY OF FORMATION (FRACTION) ====
'
+K=1/* 0.035
'
' ---- POROSITY OF PROPPED FRACTURED GRID BLOCKS (FRACTION) ----
'
I=1/1 J=15/29 K=1/* 0.37
I=14/14 J=15/29 K=1/* 0.37
I=27/27 J=15/29 K=1/* 0.37
I=40/40 J=15/29 K=1/* 0.37
I=53/53 J=15/29 K=1/* 0.37
I=66/66 J=15/29 K=1/* 0.37
I=79/79 J=15/29 K=1/* 0.37
I=92/92 J=15/29 K=1/* 0.37
'
' ---- ELEMENT OF SYMMETRY - POROSITY MULTIPLIER ----
'
MULT I=1/1 J=1/* K=1/* 0.5
END
'
' ==== PERMEABILITY (HORIZONTAL & VERTICAL) OF FORMATION (mD) ====
'
+K=1/* 0.015 0.0015
'
' ---- PERMEABILITY MULTIPLIERS OF SRV REGION (ONE WING-120X300X100 ft3)-INITIAL
PERMEABILITY MULITPLIER(IPM) ----
' HORZ VERT
MULT I=1/6 J=5/39 K=1/* 50.0 50.0
MULT I=9/19 J=5/39 K=1/* 50.0 50.0
MULT I=22/32 J=5/39 K=1/* 50.0 50.0

182
MULT I=35/45 J=5/39 K=1/* 50.0 50.0
MULT I=48/58 J=5/39 K=1/* 50.0 50.0
MULT I=61/71 J=5/39 K=1/* 50.0 50.0
MULT I=74/84 J=5/39 K=1/* 50.0 50.0
MULT I=87/97 J=5/39 K=1/* 50.0 50.0
END
'
' ==== TRANSMISSIBILITY MODIFIERS FOR FRACTUREDE BLOCKS ====
' Tx TY TZ
I=1/1 J=15/29 K=1/* 1.0 27.0 261.0
I=14/14 J=15/29 K=1/* 1.0 27.0 261.0
I=27/27 J=15/29 K=1/* 1.0 27.0 261.0
I=40/40 J=15/29 K=1/* 1.0 27.0 261.0
I=53/53 J=15/29 K=1/* 1.0 27.0 261.0
I=66/66 J=15/29 K=1/* 1.0 27.0 261.0
I=79/79 J=15/29 K=1/* 1.0 27.0 261.0
I=92/92 J=15/29 K=1/* 1.0 27.0 261.0
'
' ---- ELEMENT OF SYMMETRY - POROSITY MULTIPLIER ----
'
MULT I=1/1 J=1/* K=1/* 1.0 0.5 0.5
'
' ---- PRESSURE DEPENDENT PERMEABILITY MULITPLIERS INSIDE SRV - VARIABLE MEAN TOTAL STRESS CASE ----
'
STRMULT STRESS PRODUCTION
'
I=1/6 J=5/14 K=1/*
I=1/6 J=30/39 K=1/*
I=2/6 J=15/29 K=1/*
I=9/19 J=5/14 K=1/*
I=9/19 J=30/39 K=1/*
I=9/13 J=15/29 K=1/*
I=15/19 J=15/29 K=1/*
I=22/32 J=5/14 K=1/*
I=22/32 J=30/39 K=1/*
I=22/26 J=15/29 K=1/*
I=28/32 J=15/29 K=1/*
I=35/45 J=5/14 K=1/*
I=35/45 J=30/39 K=1/*
I=35/39 J=15/29 K=1/*
I=41/45 J=15/29 K=1/*
I=48/58 J=5/14 K=1/*
I=48/58 J=30/39 K=1/*
I=48/52 J=15/29 K=1/*
I=54/58 J=15/29 K=1/*
I=61/71 J=5/14 K=1/*
I=61/71 J=30/39 K=1/*
I=61/65 J=15/29 K=1/*
I=67/71 J=15/29 K=1/*
I=74/84 J=5/14 K=1/*
I=74/84 J=30/39 K=1/*
I=74/78 J=15/29 K=1/*
I=80/84 J=15/29 K=1/*
I=87/97 J=5/14 K=1/*
I=87/97 J=30/39 K=1/*
I=87/91 J=15/29 K=1/*
I=93/97 J=15/29 K=1/*
'
' ---- STRESS FACTOR (S) = 8.4 STARTING FROM 2000 PSI ----
'

183
' Effective Kx Ky Kz
' Stress
3688 1.00000000 1.00000000 1.00000000
3731 0.87880844 0.87880844 0.87880844
3774 0.76903837 0.76903837 0.76903837
3817 0.66989236 0.66989236 0.66989236
3859 0.58061636 0.58061636 0.58061636
3902 0.50049720 0.50049720 0.50049720
3945 0.42886013 0.42886013 0.42886013
3988 0.36506659 0.36506659 0.36506659
4031 0.30851214 0.30851214 0.30851214
4074 0.25862450 0.25862450 0.25862450
4117 0.21486170 0.21486170 0.21486170
4159 0.17671037 0.17671037 0.17671037
4202 0.14368419 0.14368419 0.14368419
4245 0.11532228 0.11532228 0.11532228
4288 0.09118791 0.09118791 0.09118791
4331 0.07086710 0.07086710 0.07086710
4374 0.05396740 0.05396740 0.05396740
4417 0.04011671 0.04011671 0.04011671
4459 0.02896224 0.02896224 0.02896224
4502 0.02016941 0.02016941 0.02016941
4545 0.01342093 0.01342093 0.01342093
'
' ---- PRESSURE DEPENDENT PERMEABILITY MULITPLIERS OUTSIDE SRV - VARIABLE MEAN TOTAL STRESS CASE ----
'
STRMULT STRESS PRODUCTION
'
I=1/* J=1/4 K=1/*
I=1/* J=40/* K=1/*
I=7/8 J=5/39 K=1/*
I=20/21 J=5/39 K=1/*
I=33/34 J=5/39 K=1/*
I=46/47 J=5/39 K=1/*
I=59/60 J=5/39 K=1/*
I=72/73 J=5/39 K=1/*
I=85/86 J=5/39 K=1/*
I=98/* J=5/39 K=1/*
'
' ---- STRESS FACTOR (S) = 6.0 STARTING FROM 2000 PSI ----
'
' Effective Kx Ky Kz
' Stress
3688 1.00000000 1.00000000 1.00000000
3731 0.91237313 0.91237313 0.91237313
3774 0.83093231 0.83093231 0.83093231
3817 0.75532405 0.75532405 0.75532405
3859 0.68521320 0.68521320 0.68521320
3902 0.62028185 0.62028185 0.62028185
3945 0.56022833 0.56022833 0.56022833
3988 0.50476632 0.50476632 0.50476632
4031 0.45362397 0.45362397 0.45362397
4074 0.40654308 0.40654308 0.40654308
4117 0.36327836 0.36327836 0.36327836
4159 0.32359670 0.32359670 0.32359670
4202 0.28727653 0.28727653 0.28727653
4245 0.25410718 0.25410718 0.25410718
4288 0.22388828 0.22388828 0.22388828
4331 0.19642926 0.19642926 0.19642926
4374 0.17154876 0.17154876 0.17154876

184
4417 0.14907421 0.14907421 0.14907421
4459 0.12884134 0.12884134 0.12884134
4502 0.11069375 0.11069375 0.11069375
4545 0.09448251 0.09448251 0.09448251
END
'
' ==== INTIAL CONDITIONS (INIAILIZATION) ====
'
' INITIAL PRESSURE
END
' GAS AND WATER SATURATION
END
PREF=2000. DREF=5300. DGOC=15000. DWOC=15000.
END
' ====================================================================================
' PVT DATA
' ====================================================================================
' PRESSURE EGT VISGT
' (PSI) (SCF/RB) (CP)
14.7 4.4892 0.01352
147 45.317 0.01358
279 86.898 0.01368
411 129.19 0.01380
676 215.82 0.01409
808 260.05 0.01426
1073 350.03 0.01465
1205 395.63 0.01486
1470 487.59 0.01534
1602 533.74 0.01559
1866 625.81 0.01614
2028 681.59 0.01650
2205 741.73 0.01691
2557 858.74 0.01777
2734 915.18 0.01821
2910 970.02 0.01866
3087 1023.1 0.01912
3248 1070.2 0.01955
3513 1143.9 0.02025
3627 1174.3 0.02054
4000 1268.5 0.02152
8000 1916.5 0.03056
12000 2255.7 0.03743
16000 2481.2 0.04311
20000 2649.7 0.04805
END
' ====================================================================================
' PHYSICAL PROPERTIES DATA
' ====================================================================================
' ==== DENSITIES OF OIL-GAS-WATER: (LBM/FT**3) ====
'
DENO= 52.10 DENG= 0.04688 DENW= 62.4
'
' ==== COMPRESSIBILITIES OF FORMATION-OIL-WATER: (ONE/PSI) ====
'
CFOR=3.500E-06 CO=8.38E-06 CW=3.10E-06
'
' ==== WATER INPUT FORM VOL FACT; REF PRES (FORM-WATER): (PSI ====
'
BWI= 1.026 PFOR= 500.0 PW= 500.0
'

185
' ==== VISCOSITY OF WATER (CP) VISCOSITY PRESSURE VARIANCE: (CP/PSI) ====
'
VWI= 0.25
'
' ==== OIL VISCOSITY PRESSURE VARIANCE ABOVE THE BUBBLE POINT: (CP/PSI) ====
'
CVO=8.72D-05 ' =(VO2-VO1)/(P2-P1)
'
' ==== WATER VISCOSITY PRESSURE VARIANCE: (CP/PSI) ====
'
CVW=0.0000000 ' NORMALLY SET TO ZERO
'
=====================================================================================
' RELATIVE PERMEABILITY & CAPILLARY PRESSURE
' ====================================================================================
' GAS RELATIVE PERMEABILITY CURVE FITTING QUADRATIC
GQUAD=OFF ' NORMALLY OFF
' SW KRW KRG PCOG
' ==== ==== ==== ====
0.25 0 1.000 1000.3
0.3 0.007 0.793 271.80
0.35 0.016 0.69 110.54
0.4 0.029 0.608 60.298
0.45 0.047 0.52 38.098
0.5 0.072 0.453 26.232
0.55 0.105 0.38 19.377
0.6 0.148 0.320 14.533
0.65 0.201 0.26 9.6890
0.7 0.266 0.205 4.8404
0.8 0.438 0.111 4.4760
0.9 0.679 0.039 3.9670
1 1.000 0.000 0.0000
END
' ====================================================================================
' RECURRENT DATA
' ====================================================================================
'
TIME= 15/06/2006 START
'
' ====================================================================================
' SIMULATOR CONTROLS
' ====================================================================================
' ==== CONVERGENCE TOLERENCE ====
'
RTOL 0.01 ' RESIDUAL COVERGENCE TOLERANCE (RANGE .0001 TO .5)
PTOL .100000E-00 ' PRESSURE COVERGENCE TOLERANCE (RANGE .1 TO 100.)
STOL .500000E-02 ' SATURATION CONVERGENCE TOLERANCE (RANGE .0001 TO .5)
'
' ==== AUTOMATIC TIME STEP CONTROLS ====
'
PNORM= 500.00 ' MAX GRID BLOCK PRESSURE CHANGE (PSI/KPA)
BPNORM= 500.00 ' MAX GRID BLOCK BUBBLE POINT PRESSURE CHANGE (PSI/KPA)
SNORM= 0.2 ' MAX GRID BLOCK SATURATION CHANGE (FRACTION)
'
' ==== TIME STEP CONTROLS ====
'
TS = 0.5 ' FIRST TIMESTEP SIZE (DAYS)
M-TS = 10.0 ' MAXIMUM TIMESTEP SIZE (DAYS)
N-TS = 99999 ' MAXIMUM NUMBER OF TIMESTEPS (SET TO LARGE VALUE)
'

186
' ==== IMPLICIT SWITCHING CONTROLS (THRESHOLDS) ====
'
PNIMP= 2000 ' BUBBLE POINT THRESHOLD (USUALLY 1/4 OF BPNORM)
SNIMP= 0.1 ' SATURATION THRESHOLD (USUALLY 1/4 OF SNORM)
'
' ==== SOLUTION CONTROLS ====
'
IMPLICIT ' FORMULATION (RE) SETTING: IMPLICIT EXPLICIT OFF (START WITH EXPLICIT)
NEIGH=OFF ' IMPL/EXPL WELL CELLS: NEIGH= ON PLUS (USUALLY OFF)
INFO= OFF ' GENERAL OUTPUT DETAIL: INFO= ON OFF PLUS (USUALLY OFF)
'
' ==== OUTPUT CONTROLS ====
'
R-OUT=ON
W-OUT=ON
RESTART=OFF
FORECAST=1
CPLOT=ON
PR+W PR+G PR+P PR-B PR+V PR+K PR+C
W-LONG
'
' ====================================================================================
' WELL DESCRIPTION
' ====================================================================================
WELL=PROD START NEW
PRODUCER
DIREC=1
WR= 0.2500 CC=0.37000 FF=1.000000 SKIN= 0.0000
' HORIZONTAL WELL COMPLETION - PERFORATION LOCATIONS
LAYER=1 I=1 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
LAYER=2 I=14 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
LAYER=3 I=27 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
LAYER=4 I=40 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
LAYER=5 I=53 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
LAYER=6 I=66 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
LAYER=7 I=79 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
LAYER=8 I=92 J=22 K=9 GRID=1 LMULT = 1000.0 HDOFF= 0.0
'
CONSTRAINT = 1 GAS = 7000000.00000 ' SCF/DAY
CONSTRAINT = 2 PRESS = 300.0000 Min ' PSI
'
LIST=SWITCH
BACK=OFF
WCUT= OFF
'
' ==== TIME AND ACTION ====
TIME= 1.0 CONTINUE
TIME= 4.0 CONTINUE
TIME= 15.0 CONTINUE
TIME= 30.0 CONTINUE
TIME= 60.0 CONTINUE
TIME= 85.0 CONTINUE
TIME= 166.0 CONTINUE
TIME= 365.0 CONTINUE
TIME= 547.5 CONTINUE
TIME= 588.5 CONTINUE
TIME= 730.0 CONTINUE
TIME= 912.5 CONTINUE
TIME= 1095.0 CONTINUE
TIME= 1296.0 STOP

187
Case – 6.6.fem : The Input File for Stress - Strain Model
Stress - Strain Model - Case 6.6
Metric units, 3D , 1 material, linearly elastic
COM ( NOTE: above TITLE must be 2 lines )
COM
COM ***********************************
COM * Taurus Reservoir Solutions Ltd. *
COM * FEM3D *
COM * 3-D STRESS STRAIN MODEL *
COM * VERSION 3.11 *
COM * 06-DECEMBER-2004 *
COM ***********************************
COM
COM RECORD B
COM
COM --- IUNIT,INPUT UNITS : 0 = metric (kPA, m, deg C)
COM | 1 = English (psia, ft, deg F)
COM |
COM |
COM V
COM *----*
1
COM *----*
COM RECORD C
COM
COM ,--Ngrids TOTAL NUMBER OF GRIDS AND SUB-GRIDS
COM | ,--REFERENCE DEPTH TO THE TOP OF GRID
COM | | (depth is negative downward)
COM | | ,--A_type, analysis geometry type:
COM | | | 0=3D_Solid, 1=Radial
COM | | | ,--Mat_cnt, NUMBER OF MATERIALS
COM | | | | ,--Mat_hyst, STRESS-STRAIN HYSTERESIS
COM | | | | | 0=no, 1=yes (all mat.types)
COM | | | | | ,--Iter_flg, SOLVER OPTIONS:
COM | | | | | | 0=direct solver;1=ILU iterative solver
COM | | | | | | 3=CG solver
COM | | | | | | ,--Th_comp_flg, Thermal compaction
COM | | | | | | | and hardening option(uses Temp_ref
COM | | | | | | | of .gii file in calcs):
COM | | | | | | | 0=no; 1=thermal compaction
COM | | | | | | | ,--Read_flg, Grid read option:
COM | | | | | | | | 0=read dx,dy,dz data;
COM | | | | | | | | 1,2=read CPG coord by nodes
COM | | | | | | | | 1=reserv.grid only
COM | | | | | | | | 2=read total grid
COM | | | | | | | | 3=read faulted FEM grid by
COM | | | | | | | | FaultPrep preprocessor
COM | | | | | | | | ,--Bore Radius for Radial analyses
COM | | | | | | | | | ,---G_type, grid type
COM | | | | | | | | | | 1=Continuous/Refined,
COM | | | | | | | | | | 2= Faulted
COM | V | | | | | | V |
COM V m/ft V V V V V V m/ft V
COM *---------|----*----*----*----*----*----*----*----*
1 -5250 0 1 0 1 0 0 0 1
COM *---------|----*----*----*----*----*----*----*----*
COM
COM RECORD C1
COM

188
COM Read always - PRIMARY GRID SIZE
COM ,--Nx,NUMBER OF Primary GRID BLOCKS IN X DIRECTION
COM | ,--Ny,NUMBER OF Primary GRID BLOCKS IN Y DIRECTION
COM | | ,--Nz,NUMBER OF PRIMARY GRID BLOCKS IN Z DIRECTION
COM | | |
COM | | |
COM V V V
COM *----*----*
110 43 40
COM *----*----*
COM RECORD D
COM
COM Defines relative position of the reservoir grid in the
COM primary FEM grid
COM ,--N_low_dx, NUMBER OF X-GRID BLOCKS BEFORE PRIMARY RES.GRID
COM | ,--N_low_dy,NUMBER OF Y-GRID BLOCKS BERORE PRIMARY RES. GRID
COM | | ,--N_low_dz,NUMBER Z-GRID BLOCKS BEFORE PRIMARY RES.GRID
COM | | |
COM V V V
COM *----*----*
0 0 13
COM *----*----*
COM
COM ,--N_high_dx, NUMBER OF X-GRID BLOCKS AFTER RES.GRID
COM | ,--N_high_dy,NUMBER OF Y-GRID BLOCKS AFTER RES. GRID
COM | | ,--N_high_dz,NUMBER Z-GRID BLOCKS AFTER RES.GRID
COM | | |
COM V V V
COM *----*----*
0 0 10
COM *----*----*
COM RECORD E
COM
COM *** PRIMARY GRID SIZE DATA ***
COM ( read the following 3 lines only if Read_flg = 0)
COM ( otherwise no data )
COM ---- ARRAY OF VALUES SPECIFYING LENGTHS OF GRID BLOCKS IN X DIRECTION - (Nx values ) ----
COM <---------------------------- m/ft --------------------------------------->
COM
2 3 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10
10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30 30 30 10 10 5 4 2 4 5 10 10 30 30
30 30 10 10 5 4 2 4 5 10 10 30 30 150 150 350 500 500 1000 1000 1000 1000 1000 1000 1000
COM
COM ---- ARRAY OF VALUES SPECIFYING LENGTHS OF GRID BLOCKS IN Y DIRECTION - (Ny values ) ----
COM <---------------------------- m/ft --------------------------------------->
COM
1000 1000 500 200 100 60 22 22 18 12 10 10 10 8 5 4 4 4 4 4 2 2 2 4 4 4 4 4 5 8 10 10 10 12 18 22 22 60 100
200 500 1000 1000
COM
COM ---- ARRAY OF VALUES - LENGTHS OF GRID IN Z DIRECTION - (Nz values, from top down ) ----
COM <---------------------------- m/ft --------------------------------------->
COM
1000 1000 1000 500 500 300 200 200 100 100 40 30 30 14 10 10 5 4 2 2 2 2 2 2 2 4 5 10 10 14 30 30 40 100 100
200 200 300 500 500
COM
COM *** GROUP I - STRESS INITIALIZATION ***
COM
COM RECORD I0
COM
COM ,--FEM_ref_Depth: REFERENCE DEPTH FOR STRESS INITIALIZATION

189
COM | (z is negative downward)
COM | ,--Mode_instr: MODE OF STRESS INITIALIZATION
COM | | 0 - read stresses at FEM_ref_depth + gradients
COM | | 1 - read stresses by layers
COM | | 2 - initialization by variable gradient (using rock, fluid
COM V | saturations and K0 stress ratio) (project specific)
COM m/ft V
COM -----|----*
5300 0
COM -----|----*
COM
COM **** UNIFORM STRESS INITIALIZATION ****
COM Read this data only if Mode_instr = 0 was specified
COM
COM RECORD I1
COM
COM ,--FEM_Ref_Sx: INITIAL TOTAL STRESS IN X-DIR (compression= +)
COM | ,--FEM_Ref_Sy: INITIAL TOTAL STRESS IN Y-DIR
COM | | ,--FEM_Ref_Sz: INITIAL TOTAL STRESS IN Z-DIR
COM | | |
COM V V V
COM ----- kPa / psia ----
COM -----|---------|---------|
4505 5989 6572
COM -----|---------|---------|
COM
COM RECORD I2
COM
COM ,--FEM_Grad_Sx:INITIAL TOTAL X-STRESS GRADIENT WITH DEPTH (increase down=+)
COM | ,--FEM_Grad_Sy: INITIAL TOTAL Y-STRESS GRDIENT WITH DEPTH
COM | | ,--FEM_Grad_Sz:INITIAL TOTAL Z-STRESS GRADIENT WITH DEPTH
COM | | |
COM V V V
COM --- kPa/m / psia/ft ---
COM -----|---------|---------|
0.0 0.0 0.0
COM -----|---------|---------|
COM
COM RECORD F
COM
COM *** STATIC/DYNAMIC LOAD/B.C. EVENT CONTROL ***
COM
COM (Notice: have at least one load event in order to read in fixity B.C.)
COM
COM *** STATIC AND/OR DYNAMIC LOADS ***
COM Read the number of static and transient load events
COM ,--num_static: Number of Quasi-Static (initialization) Load Events
COM | ,--Indic_transient: Indicator for transient solution:
COM | | 0=initialization only (do not read transient loads)
COM | | 1=do transient stress solve, read transient loads
COM V V
COM *----*
0 1
COM *----*
COM *** GROUP F2 - TRANSIENT (DYNAMIC SIMULATION) LOAD EVENT DATA ***
COM (Read if Indic_transient >0)
COM
COM RECORD F2 - Transient Load Event Data
COM
COM ---- ZDispfl option to zero the displacement array at the beginning ----

190
COM | of the transient coupled simulation (1=zero , 0=retain)
COM | ,--Num_Disp_BC: number of generalized displacement BC sets
COM | | ,--Num_Face_Loads: number of face regions with normal forces
COM | | | ,N_Tab_Loads: number of transient load tables
COM | | | | ,N_frac_loads: Number of dynamic/static frac loaded regions
COM | | | | | (+ dynamic loading; - static loading)
COM | | | | | | ,--Special fixity key (0: no additional fixities; 1: need file xx.fix
COM | | | | | | | (form: total # of fixed nodes, I,J,K,Fi,Fy,Fz)
COM V V V V V V V
COM *----*----*----*----*----*----*
1 0 0 0 0 0 0
COM *----*----*----*----*----*----*
COM
COM RECORD F2A
COM
COM DISPLACEMENT FIXITY KEYS FOR SIDES (FACES) OF THE MODEL :
COM
COM INPUT 6 LINES FOR SIDES I=1,..,6 IN THE FOLLOWING ORDER:
COM I=1 - TOP (z = Lz)
COM I=2 - BOTTOM (z = 0)
COM I=3 - LEFT (y = 0)
COM I=4 - RIGHT (y = Ly)
COM I=5 - FRONT (x = Lx)
COM I=6 - BACK (x = 0)
COM
COM ,--Anl_fix_key(1,I): x-direction displacement on face I:0=free,1=fixed
COM | ,--Anl_fix_key(2,I): y-direction displacement: 0=free, 1=fixed
COM | | ,--Anl_fix_key(3,I): z-direction displacement: 0=free, 1=fixed
COM | | |
COM V V V
COM *----*----*
0 0 0
0 0 1
0 1 0
0 1 0
1 0 0
1 0 0
COM *----*----*
COM
COM *** GROUP M - MATERIAL PROPERTIES (CONSTITUTIVE MODELS) ***
COM
COM **** MATERIAL PROPERTIES ****
COM Read as many sets of data as there is material types,total of Mat_cnt
COM Each set consists of record M0 plus one of the Records M1,M2 or M3
COM
COM RECORD M0
COM
COM ,--Mat_type: type of material (1=linearly elastic, 2=hyperbolic, 3=tables,
COM | 4=hypo-plastic, 6=Barton-Bandis, 11=lin.el./EP
COM | ,--KEY1 (=ISIGE for Mat_type=2,
COM | | =independent variable for tables for Mat_type=3:
COM | | 0= min eff stress
COM | | 1= mean eff stress
COM | | 2= depth
COM | | ,--KEY2 (=ISIGB for Mat_type=2)
COM | | | ,--KEY3 (=IDILATE for Mat_type=2, OPTION FOR MODELING DILATION:
COM | | | | 0= NO DILATION
COM | | | | 1= SHEAR-INDUCED DILATION
COM | | | | 2= STRESS-RATIO-INDUCED DILATION
COM V V V V

191
COM *----*----*----*
1 0 0 0
COM *----*----*----*
COM
COM RECORD M1
COM
COM Input for Mat_type = 1 - linearly elastic material
COM --------------------------------------------------
COM
COM ,--Mat_E: YOUNG'S MODULUS
COM | ,--Mat_nu: POISSON RATIO
COM | | ,--Mat_cte: COEFF. OF THERMAL EXPANSION (LINEAR)
COM | | | ,--Mat_ke: MATRIX (GRAIN) MODULUS (1/compress.)
COM | | | | ,--Mat_Density: DENSITY (mass/volume)
COM | | | | |
COM | | | | |
COM V V V V V
COM kPa - 1/deg C kPa kg/m3
COM psia - 1/deg F psia lb/cuft
COM -----|---------|---------|---------|---------|
7.9E6 0.125 0.0 1.0E12 160
COM -----|---------|---------|---------|---------|
COM
COM *** GROUP S - SOLUTION CONTROLS ***
COM
COM RECORD S1
COM
COM **** FEM Solution Strategy ****
COM
COM a) If all Material types are elastic (Material type 1,2,3,4 or 6)
COM read the following data:
COM
COM In this case iterations can be used within time step to
COM solve problems with dilation
COM
COM ,--Solution_Type: linear/nonlinear elastic solution strategy
COM | Set to 1 - Constant K matrix during time step
COM | ,--Max_Iterations:
COM | | 0 = no iteration during time/load step (for linear and
COM | | nonlinear elasticity without dilation)
COM | | n > 0 = use max of n iterations during solution to
COM | | eliminate out of balance force caused by dilation
COM v v
COM *----*
1 0
COM *----*
COM RECORD S2
COM
COM ITERATIVE SOLVER PARAMETERS (Read only if Iter_flg = 1)
COM
COM a) Iterative solver specification flags (options)
COM
COM ,--Iter_iadd: SPD flag - To add amount to the main diagonal to force
COM | positive definitness (if problem with diag. dominance)
COM | 0 = no, >0 = number of adds
COM | ,--Iter_idrop: ILU flag - Type of ILU decomposition
COM | | 0 = level-based ILU (fixed fill pattern, no dropping)
COM | | 1 = drop tolerance ILU (drop terms based on the orig.
COM | | diagonal test
COM | | 2 = drop tolerance ILU, updated diag. test (preferred)

192
COM | | ,--Iter_level: Number of extra diagonals used for ILU
COM | | | (applies only for fixed fill option, Iter_drop = 0)
COM | | |
COM V V V
COM *----|----|
0 0 0
COM *----|----|
COM
COM b) Iterative solver specification data
COM
COM ,--Riter_dgadd: Amount added to the main diagonal
COM | (applies only if SPD flag > 0,i.e., Iter_idgadd >0 )
COM | Value must be > 0 if applicable.
COM | ,--Riter_dropt: Drop tolerance value for Drop ILU options
COM | | (applies only if ILU flag = 1 or 2, i.e., Iter_idrop=1,2)
COM | | Value must be > 0 if applicable.
COM | | ,--Riter_dgtol: Diagonal tolerance for checking if matrix
COM | | | is diag. dominant (enter 0.0)
COM | | | ,--Riter_ctol: Absolute convergence tolerance
COM | | | | (for each component of the solution)
COM | | | | ,--Riter_reduce: Residual norm
COM | | | | | reduction factor from the initial
COM | | | | | value
COM | | | | |
COM | | | | | NOTE: Soln is converged when the
COM | | | | | FIRST tolerance is satisfied
COM V V V V V
COM *----|---------|---------|---------|---------|
0 0 0 1.0d-20 1.0d-20
COM *----|---------|---------|---------|---------|
COM
END
EOF

Case – 6.6.gii : The Input Interface


COM **************************************************
COM * Taurus Reservoir Solutions Ltd. *
COM * GEOINT *
COM * INTERFACE BETWEEN RES AND STRESS MODEL *
COM * VERSION 3.14 *
COM * 26-APRIL-2007 *
COM **************************************************
COM RECORD A
COM
COM ,--IUNIT,INPUT UNITS : 0 = metric (kPA, m, deg C)
COM | 1 = English (psia, ft, deg F)
COM | ,--ASIZE, size of memory for arrays (integer words)
COM | | ,--ASOLVERSIZE, size of solver memory (integer words)
COM | | |
COM V V V
COM *---------*---------*
1 50e9 30e9
COM *---------*---------*
COM RECORD B
COM
COM **** NUMERICAL CONTROLS OF THE COUPLING ****
COM
COM ,--Ptol, PRESSURE DP TOLERANCE FOR CONVERGENCE
COM | (global iter between reserv. and stress model)

193
COM | ,--Accmax, MAX. ACCELERATION FACTOR ( > 1)
COM | | ,--N_noacc, NO OF T.S. WITHOUT
COM | | | ACCELERATION AFTER A RATE
COM | | | CHANGE IN THE HOST MODEL
COM | | | ,--N_solve, frequency of stress
COM | | | | soln in EXPLICIT mode:
COM | | | | 0= every t.s.
COM | | | | n= every nth t.s.
COM | | | | -1= every TCHG
COM | | | | -n= every nth t.s.+TCHG
COM | | | | ,--Key_resm, no of
COM | | | | | regions for res.perm.
COM | | | | | modifiers >=0
COM | | | | | ,--Key_resm_p, no of
COM | | | | | | regions for res.poros.
COM | | | | | | modifiers >=0
COM | | | | | | ,--Igappr, geom. iter. method (1,2,3) ...
COM | | | | | | | 0,1 = standard (Settari+Mourits)
COM | | | | | | | 2 = modified standard
COM | | | | | | | 3 D.Tran
COM V | | | | | | ,--Factor1, constr.type constant
COM kPa | | | | | | | for igappr=3
COM psia V V V V V V V
COM -----|---------|----*----*----*----*----*---------|
1.0 1.0 10 0 0 0 0 0 .0
COM -----|---------|----*----*----*----*----*---------|
COM
COM RECORD B1
COM
COM **** INITIALIZATION DATA - OUTSIDE OF RESERVOIR GRID, NULL CELLS ****
COM **** NORMALIZING STRESS TABLES ****
COM
COM ,--Press_ref, REFERENCE PRESSURE at depth=0 (used to set initial pressure
COM | in null cells of reservoir grid)
COM | ,--Temp_ref, REFERENCE TEMPERATURE (used for thermal
COM | | compaction in FEM3D)
COM | | ,--wGrad, PORE FLUID GRADIENT to set pressure in null cells
COM | | | p(m) = depth(m)*wGrad
COM | | | ,--Norm_perm, if=1 normalize stress dependent
COM | | | | permeability in Record J to initial stress
COM | | | | ,--Norm_por, if=1 normalize stress
COM V V V | | dependent porosity in Record K to
COM kpa deg C kPa/m | | initial stress (not yet implemented)
COM psia deg F psia/ft V V
COM -----|---------|---------|----*----*
0.0 0.0 0.0
COM -----|---------|---------|----*----*
COM
COM RECORD C
COM
COM ***** PRINTOUT OPTIONS AND CONTROLS *****
COM
COM ,--IOUTPR: type of printout of FEM solution:
COM | 0 - use FEM print routines (in element order)
COM | 1 - print 3-D arrays of variables consistent with host grid (read 16 keys of stresses, displacements...)
COM | 2 - Same as IOUTPR=1, and read additional keys to print mechanical properties
COM |
COM | ,--IOUTCA: controls postprocessing files for TERAPRO/SEISMIC:
COM | | 0 - no files
COM | | 1 - write TERAPRO data to .gc* with freq. of arrays

194
COM | | +-2 - write SEISMIC data to .seis with freq. of arrays
COM | | +-3 - write BOTH with freq. of arrays
COM | | +2 or +3 means write reservoir arrays into .seis over FEM grid
COM | | -2 or -3 means write reservoir arrays into .seis only over res grid
COM | | ,--idisp_avg: Method for averaging z-displacement for TERAPRO:
COM | | | 0 - compute block center values
COM | | | 1 - compute top of block values
COM | | | ,--Key_var_seis: control on content of SEIS file
COM | | | | 0 - Heriot Watt set of arrays (Appendix E)
COM | | | | 1 - as 0 plus Sz and strainz (V. Sen)
COM | | | | 2 - as 1 plus vert.displ (averaged according to idisp_avg)
COM | | | | 3 - as 2 but suppress writing reservoir arrays
COM | | | |
COM V V V V
COM *----*----*----*
1 1 1 0
COM *----*----*----*
COM
COM Read these 2 print controls (Records D and E) ONLY if IOUTPR > 0
COM
COM RECORD D
COM
COM ,--IPLNPR: choice of 2-D planes to print 3-D arrays:
COM | 0 - x-y (areal), 1 - x-z (crossections)
COM V
COM *----*
0 0
COM *----*
COM
COM RECORD E
COM
COM KEYPRT(1-16) keys to print arrays:
COM 0=no, 1=yes (total stresses), 2=yes (eff. stresses)
COM
COM ,--print stress Sx
COM | ,--print stress Sy
COM | | ,--print stress Sz
COM | | | ,--print shear stress Txy
COM | | | | ,--print shear stress Txz
COM | | | | | ,--print shear stress Tyz
COM | | | | | | ,--principal max. stress Smax
COM | | | | | | | ,--principal min. stress Smin
COM | | | | | | | | ,--Displacement Ux
COM | | | | | | | | | ,--Displacement Uy
COM | | | | | | | | | | ,--Displacement Uz
COM | | | | | | | | | | | ,--Hysteresis keys
COM | | | | | | | | | | | | ,--Pressure
COM | | | | | | | | | | | | | ,--Stress level
COM | | | | | | | | | | | | | | ,--Volume.
COM | | | | | | | | | | | | | | | strain
COM | | | | | | | | | | | | | | | Savg
COM V V V V V V V V V V V V V V V V
COM *----*----*----*----*----*----*----*----*----*----*----*----*----*----*----*
1 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1
COM *----*----*----*----*----*----*----*----*----*----*----*----*----*----*----*
COM RECORD F
COM
COM **** MONITORING AND DEBUG PRINTOUT - Use with caution !!! ****
COM
COM ,--Monitor, PRINT CONTROL:

195
COM | 0 = NO EXTRA PRINTOUT
COM | 1 = SOME DEBUG FOR FIRST 6 NODES
COM | 2 = 1 + POROSITIES
COM | 3 = 2 + STRESSES AND STRAINS AFTER EACH ITER.
COM | ,--Mode_prt, ARRAY PRINT CONTROL MODE:
COM | | 0 = CONTROLLED BY HOST MODEL ARRAY FREQUENCY
COM | | 1 = INDEPENDENT, BASED ON IPRTSTR FREQUENCY
COM | | 2 = PRINT BOTH AT HOST FREQUENCY AND IPRTSTR FREQUENCY
COM | | ,--IPRTSTR, FREQUENCY OF PRINTING STRAIN/STRESS ARRAYS
COM | | | (used if Mode_prt = 1 or 2)
COM | | | ,--Monit_sol, KEY TO MONITOR PERFORMANCE OF FEM SOLVERS:
COM | | | | 0 = no print
COM | | | | 1 = print timings and other info
COM | | | |
COM V V V V
COM *----*----*----*
0 0 0 0
COM *----*----*----*
COM RECORD G
COM
COM **** PRINT CONTROLS FOR MONITORING STRESSES ALONG A WELLBORE ****
COM
COM ,--Imonit_str: output level:
COM | 0 - none, 1 - t.s.summary, 2 - as 1 + stresses+p
COM | ,--IPLT_STR: postprocessing plot/link files
COM | | 0=no
COM | | 1=write .trs (p,T+stresses) and .trd (displacements) files
COM | | for Mmonit_str elements (blocks) specified in Record I
COM | | (external file names xxxx.trsn and xxxx.trdn where
COM | | n is the number of the block in the input order)
COM | | 2=as =1+write fort57 with link data for stand alone FEM3D)
COM | | ,--Mmonit_str: number of blocks for which the postprocessing
COM | | | files .trs and .trp are written (2 files/block, max=10)
COM | | | ,--IPRT_RESET: reset print controls after fracture
COM | | | | re-initiation: 0 = no, 1 = print arrays
COM | | | |
COM V V V V
COM *----*----*----*
0 1 0 0
COM *----*----*----*
COM RECORD H
COM
COM **** REGION TO SEARCH FOR MINIMUM STRESS ****
COM (should define a line within the grid)
COM
COM ,--Im1 - first I-index (in FEM grid)
COM | ,--Im2 - last I-index (in res. indexing over FEM grid)
COM | | ,--Jm1
COM | | | ,--Jm2
COM | | | | ,--Km1
COM | | | | | ,--Km2
COM | | | | | |
COM | | | | | |
COM V V V V V V
COM *----*----*----*----*----*
1 110 1 43 1 40
COM *----*----*----*----*----*
COM
END
EOF

196

You might also like