+ All Categories
Home > Documents > On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir...

On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir...

Date post: 05-Jul-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
234
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2016 On Hydraulic Fracturing of Tight Gas Reservoir Rock Maulianda, Belladonna Maulianda, B. (2016). On Hydraulic Fracturing of Tight Gas Reservoir Rock (Unpublished doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/27177 http://hdl.handle.net/11023/2906 doctoral 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
Transcript
Page 1: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2016

On Hydraulic Fracturing of Tight Gas Reservoir Rock

Maulianda, Belladonna

Maulianda, B. (2016). On Hydraulic Fracturing of Tight Gas Reservoir Rock (Unpublished doctoral

thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/27177

http://hdl.handle.net/11023/2906

doctoral 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

Page 2: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

UNIVERSITY OF CALGARY

On Hydraulic Fracturing of Tight Gas Reservoir Rock

by

Belladonna Maulianda

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR PHILOSOPHY IN PETROLEUM ENGINEERING

GRADUATE PROGRAM CHEMICAL AND PETROLEUM ENGINEERING

CALGARY, ALBERTA

APRIL 2016

© Belladonna Maulianda 2016

Page 3: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Abstract

Production of shale and tight gas resources is increasing which is helping to counterbalance

the conventional gas resource production decline. In 2014, shale and tight gas were 4% and 47%

of total Canadian natural gas production, respectively. By 2035, the National Energy Board

forecasts shale and tight gas production together will represent 90% of Canada’s natural gas

productions. In Canada, shale and tight gas production activities are located mainly in Western

Canada Sedimentary Basin (WCSB). The tight gas Glauconitic Formation in the Hoadley Field in

Alberta, Canada requires hydraulic fracturing of horizontal well completions because of its low

permeability of 0.07 mD. Fracture network drainage volume and enhanced permeability created

by the hydraulic fracture and the natural fracture interaction are the major enabler of commercial

production. The research documented in this thesis investigates the characteristics of the fracture

network or stimulated rock volume (SRV) caused by hydraulic fracturing. Specifically, the

dimensions of SRV, permeability, pore pressure, and in-situ stresses are examined during

hydraulic fracturing and production. Even though this topic has been examined since the early

2000s, the results provide new techniques to determine SRV properties. Three different approaches

were investigated. The first handles the impact of SRV dimensions and Young’s modulus on the

SRV effective permeability during hydraulic fracturing by using three-dimensional finite element

analysis including an investigation of fracture aperture and spacing within the SRV using a new

semi-analytical approach. The second investigates the impact of rock mechanical properties and

injected volume during hydraulic fracturing on SRV dimensions using a new analytical model.

The third explores a new nonlinear partial differential equation together with rate transient analysis

to evaluate how the SRV evolves versus distance and time with a history match of the gas flow

rate profile. The results demonstrate that the dimensions and characteristics of the hydraulic

fracture network can be estimated for the Hoadley Field.

ii

Page 4: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Acknowledgement

First, I would like to say thank you to Allah Subhanahu wa Ta’ala to give me the chance to

study at University of Calgary pursuing my PhD.

Second, I would like to say thank you to my supervisor Dr. Ian Gates to give me the opportunity

to be his student. And for his thorough support and guidance for my PhD research. His patience

and knowledge are tremendous. His guidance is helping me greatly to finish my PhD.

Third, I would like to say thank you to my co-supervisor Dr. Ron Chik-Kwong Wong for

teaching me his detailed knowledge and experience in rock mechanics. His presence has excelled

my study beyond belief. His kindness and work ethic have taught me to be a better student and

person.

Fourth, I would like to say thank you to my previous supervisor Dr. Geir Hareland who gave

me the opportunity to be in his drilling research group. And for his support to attend several

respected conferences and to meet knowledgeable industry persons.

My gratitude to Talisman, NSERC, Pason, and 7G for the research funding. My sincere thanks

to Dr. Bin Xu from Bitcan G&E and Qiang Chen for their help during my Abaqus FEA simulation

and Sandy Wang for her help in rate transient analysis.

My gratitude to Dr. Eaton and ConocoPhillips Canada especially John Henderson to provide

the Hoadley Field data. My gratitude to Weatherford Canada for their permission to use StabView.

My gratitude to Dr. Roberto Aguilera to teach me the basic of naturally fractured unconventional

reservoirs. My gratitude to Dr. Sudarshan Mehta, Dr. Jalel Azaiez and Dr. U.T. Sundararaj for

guidance throughout my difficult times. My thanks to Patricia Teichrob for her companionship

throughout difficult times.

I would like to say thank you to Dr. Sudarshan Mehta, Dr. Robert Gordon Moore, Dr. Larry

Lines, and Dr. Christopher Hawkes to be in my PhD defense committee.

Lastly, I would like to say thank you to my husband Aqsha Aqsha for his patience and support

during my study even though he is also busy finishing his PhD, my “special” son Nafi Rabbani

Aqsha for being special and cheering me up, my soon to be born baby for his/hers patience with

me, and my family/friends to support me throughout this difficult times.

iii

Page 5: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Dedication

I dedicate this PhD to my loving husband, son, soon to be born baby, and family/friends.

iv

Page 6: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table of Contents

Abstract .................................................................................................................................... ii

Acknowledgement .................................................................................................................. iii

Dedication ............................................................................................................................... iv

Table of Contents..................................................................................................................... v

List of Tables ........................................................................................................................... x

List of Figures ......................................................................................................................... xi

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

1.1 Tight Gas Sands ............................................................................................................... 1

1.2 Basic Geology of Tight Gas Sand Reservoirs .................................................................. 4

1.2.1 Continuous Gas Accumulation ................................................................................. 4

1.2.2 Conventional Gas Accumulations ............................................................................. 7

1.3 Naturally Fractured Reservoirs ...................................................................................... 11

1.3.1 Geological Classification ........................................................................................ 11

1.3.1.1 Pore System Classification ..................................................................................... 12

1.4 Hydraulic Fracturing ...................................................................................................... 14

1.4.1 Hydraulic Fracturing Basic Concepts ..................................................................... 14

1.4.2 Hydraulic Fracturing – Industrial Practice .............................................................. 14

1.5 Stimulated Rock Volume Concept and Application ...................................................... 18

1.6 Research Questions ........................................................................................................ 19

1.7 Thesis Outline ................................................................................................................ 20

1.8 References ...................................................................................................................... 21

CHAPTER 2: LITERATURE REVIEW ...................................................................................... 25

2.1 Failure Mechanics .......................................................................................................... 25

2.1.1 Tensile Failure ........................................................................................................ 25

2.1.2 Shear Failure ........................................................................................................... 26

2.2 Hydraulic Fracturing ...................................................................................................... 31

2.2.1 Hydraulic Fracturing from Vertical and Horizontal Wells ..................................... 31

2.2.2 Stress Interference due to Hydraulic Fracture Presence ......................................... 33

v

Page 7: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2.2.3 Hydraulic Fracturing on Horizontal Well Completion ........................................... 35

2.2.4 Simple Theories for Hydraulic Fracturing .............................................................. 36

2.2.5 Numerical Studies of Hydraulic Fracturing ............................................................ 38

2.3 Laboratory Studies of Hydraulic Fracturing .................................................................. 41

2.3.1. Laboratory Studies on Fractured Tight Sand and Shale Permeability ................... 41

2.3.2. Laboratory Studies on Hydraulic Fracture and Natural Fracture Interaction ......... 44

2.4 Microseismic Monitoring during Hydraulic Fracturing ................................................. 47

2.5 Stimulated Rock Volume ............................................................................................... 49

2.5.1 Stimulated Rock Volume Permeability Prediction ................................................. 49

2.5.2 Hydraulic Fracture – Natural Fracture Interaction .................................................. 50

2.5.3 Pressure Drop and Fracture Aperture Estimation ................................................... 51

2.5.4 Effect of Geomechanical Properties on the SRV .................................................... 52

2.6 Behavior of Naturally Fractured Reservoir .................................................................... 52

2.6.1 Flow Regimes for Multi-fractured Horizontal Well in a Naturally Fractured Reservoir ................................................................................................................................ 55

2.6.2 Flow Regions for Multi-fractured Horizontal Well in a Naturally Fractured Reservoir ................................................................................................................................ 58

2.6.3 Rate Transient Analysis (RTA) in a Naturally Fractured Reservoir ....................... 60

2.7 What is Missing in the Literature? ................................................................................. 61

2.8 References ...................................................................................................................... 62

CHAPTER 3: ESTIMATION OF FRACTURE CHARACTERISTIC WITHIN STIMULATED ROCK VOLUME USING FINITE ELEMENT AND SEMI-ANALYTICAL APPROACHES 69

3.1 Introduction .................................................................................................................... 69

3.1.1 Objective of Study .................................................................................................. 70

3.2 Literature Review ........................................................................................................... 71

3.2.1 Prediction of Stimulated Rock Volume Permeability ............................................. 71

3.2.2 Hydraulic Fracture - Natural Fracture Interaction .................................................. 72

3.2.3 Estimation of Pressure Drop and Fracture Aperture ............................................... 73

3.2.4 Rock Geomechanical Properties Effect on Stimulated Rock Volume .................... 73

3.3 Hoadley Field Properties ................................................................................................ 74

3.4 Finite Element Analysis Model ...................................................................................... 76

3.4.1 Constitutive Model for Tight Sand in Finite Element Analysis ............................. 76

3.4.2 Pore Fluid Flow in Finite Element Analysis ........................................................... 77

vi

Page 8: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.4.3 Finite Element Analysis Model .............................................................................. 77

3.4.4 Model Geometry ..................................................................................................... 78

3.4.5 Initial and Boundary Conditions ............................................................................. 79

3.5 Results and Discussion ................................................................................................... 80

3.5.1 Determination of Effective Permeability ................................................................ 80

3.5.2 Parametric Studies of Young’s Modulus ................................................................ 83

3.5.3 Pore Pressure and In-Situ Stresses From Finite Element Analysis......................... 85

3.5.4 Determination of Fracture Aperture Using Cubic Law .......................................... 93

3.5.5 Determination of Fracture Characteristics Using Semi-Analytical Approach ....... 95

3.6 Conclusions .................................................................................................................... 99

3.7 References .................................................................................................................... 101

CHAPTER 4: GEOMECHANICAL CONSIDERATION IN STIMULATED ROCK VOLUME DIMENSION MODELS PREDICTION DURING MULTI-STAGE HYDRAULIC FRACTURES IN HORIZONTAL WELLBORE – GLAUCONITIC TIGHT FORMATION IN HOADLEY FIELD ..................................................................................................................... 103

4.1 Introduction .................................................................................................................. 104

4.1.1 Objective of Study ................................................................................................ 105

4.2 Literature Review ......................................................................................................... 105

4.2.1 Failure Mechanics ................................................................................................. 105

4.2.1.1 Tensile Failure ...................................................................................................... 105

4.2.1.2 Shear Failure ......................................................................................................... 106

4.2.2 Stimulated Rock Volume (SRV) Models for Tight Rock Unconventional Reservoirs ............................................................................................................................ 107

4.2.3 Microseismic Monitoring during Hydraulic Fracturing ....................................... 111

4.3 Hoadley Field Project ................................................................................................... 113

4.4 Research Workflow ...................................................................................................... 118

4.4.1 Derivation of Equations for the SRV Dimensions ................................................ 118

4.4.2 Analysis of Microseismic Events.......................................................................... 120

4.4.3 Input Parameters ................................................................................................... 122

4.4.4 Calibration of Stimulated Rock Volume using Microseismic Data...................... 124

4.5 Results .......................................................................................................................... 127

4.5.1 Constant Total Stress ............................................................................................ 127

4.5.2 Total Stress Change .............................................................................................. 149

vii

Page 9: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.6 Conclusions .................................................................................................................. 154

4.7 References .................................................................................................................... 155

CHAPTER 5: DETERMINATION OF FRACTURE CHARACTERISTICS WITHIN STIMULATED ROCK VOLUME USING DIFFUSIVITY EQUATION AND PRODUCTION ANALYSIS ................................................................................................................................. 159

5.1 Introduction .................................................................................................................. 159

5.1.1 Objective of Study ................................................................................................ 160

5.2 Literature Review ......................................................................................................... 160

5.2.1 Behavior of Naturally Fractured Reservoir ........................................................... 160

5.2.2 Dual Porosity Model (Pseudo Steady State and Transient) .................................. 163

5.2.3 Flow Regimes of Multi-Fractured Horizontal Well in Naturally Fractured Reservoir .............................................................................................................................. 163

5.2.4 Flow Regions of Multi-Fractured Horizontal Well in Naturally Fractured Reservoir 164

5.3 Rate Transient Analysis (RTA) in Naturally Fractured Reservoir ............................... 165

5.3.1 RTA Concept ........................................................................................................ 165

5.4 Hoadley Field Properties .............................................................................................. 167

5.4.1 Production Data Review ....................................................................................... 167

5.5 Methodology ................................................................................................................ 168

5.5.1 Nonlinear Partial Differential Diffusivity Equation Solution ............................... 168

5.5.1.1 Nonlinear Partial Differential Diffusivity Equation For Real Gas Derivation ..... 168

5.5.1.2 Matlab Nonlinear PDE Toolbox and Code Editor ................................................ 170

5.5.1.3 Matlab Nonlinear PDE Code Editor Results ........................................................ 175

5.5.2 Application of Rate Transient Analysis Simulator (IHS Harmony Rate)............. 185

5.5.2.1 IHS Harmony Rate Transient Analysis Inputs...................................................... 185

5.5.2.2 Identifying Flow Regimes Using Type Curves .................................................... 186

5.5.2.3 Unconventional Reservoir Analysis (Unconventional Gas Module) .................... 190

5.5.2.4 History Match Using Horizontal Multifractured Enhanced Fracture Region Analytical Model ................................................................................................................. 192

5.5.2.4.1 Case 1 With Constant FCD, SRV Half-Length, and Matrix Permeability ....... 192

5.5.2.4.2 Case 2 With Constant SRV Half-Length .......................................................... 197

5.6 Conclusions .................................................................................................................. 202

5.7 References .................................................................................................................... 204

CHAPTER 6: CONCLUSIONS AND RECOMMENDATIONS .............................................. 206

viii

Page 10: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

6.1 Conclusions .................................................................................................................. 206

6.2 Recommendations ........................................................................................................ 208

Appendix A: Matlab nonlinear parabolic PDE diffusivity codes ........................................... 210

ix

Page 11: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

List of Tables

Table 3.1: Initial formation properties……………………………………………………………75

Table 3.2: SRV dimensions………………………………………………………………………75

Table 3.3: Parametric study results – effect of Young’s modulus and SRV on numbers of major and minor fractures……………………………………………………………………………….99

Table 4.1: Formation and hydraulic fracture fluid properties [39]………………………………118

Table 4.2: Glauconitic Formation properties. ............................................................................. 124

Table 4.3: Results of all stages first time step event distance, SRV growth, number of events fitted within the estimated SRV, and Figures showing results............................................................. 127

Table 4.4: The estimated SRV dimensions results and the located microseismic events. .......... 147

Table 4.5: Comparison of SRV width and hydraulic fracture port spacing................................ 147

Table 4.6: Diffusivity coefficient ratios for maximum and minimum horizontal stresses direction...................................................................................................................................................... 149

Table 4.7: Stage 7 Mohr-Coulomb failure envelope properties for Case 2. ............................... 151

Table 4.8: Stage 7 SRV dimensions and Mohr-Coulomb failure envelope properties comparison Cases 1 and 2. ............................................................................................................................. 152

Table 5.1: Input parameters……………………………………………………………………..171

Table 5.2: Matlab PDE nonlinear geometry and mesh. .............................................................. 173

Table 5.3: Matlab PDE nonlinear boundary conditions.............................................................. 174

Table 5.4: Matlab PDE nonlinear simulation time. .................................................................... 175

Table 5.5: IHS Harmony simulation input. ................................................................................. 185

Table 5.6: Case 1 simulation results for constant FCD, SRV half-length and matrix permeability...................................................................................................................................................... 194

Table 5.7: Case 2 simulation results for constant SRV half-length. ........................................... 198

x

Page 12: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

List of Figures

Fig. 1.1: Resource pyramid for natural gas [2]. .............................................................................. 2

Fig. 1.2: Tight gas production in the United States showing natural gas response as a function of DOE and GRI investment in research and State and Federal production incentives [6]. ............... 2

Fig. 1.3: Natural gas sources production by region with unconventional gas becoming the largest source of North America gas supply [7]. ........................................................................................ 3

Fig. 1.4: Tight gas resources distribution in Canada [9]. ................................................................ 4

Fig. 1.5: Geology of tight gas reservoir as a continuous gas accumulation [12]. ........................... 5

Fig. 1.6: Pressure – depth plot, Elmsworth area, Cadotte formation as an example of sub-normally pressured formation [14]. ................................................................................................................ 6

Fig. 1.7: Schematic of relative permeability curves, capillary pressures, cross section and structural map for conventional reservoir [13]. .............................................................................................. 9

Fig. 1.8: Schematic of relative permeability curves, capillary pressures, cross section and structural map for unconventional reservoir [13]. ........................................................................................ 10

Fig. 1.9: Example of orthogonal regional fractures in Devonian Antrim shale, Michigan Basin [28]........................................................................................................................................................ 11

Fig. 1.10: Chart for estimating pore-throat aperture as a function of porosities and permeability and possible ranges of oil (bpd), and gas flow rates (scfd) for different pore-throat aperture [34]........................................................................................................................................................ 12

Fig. 1.11: Porosity distribution in naturally fractured reservoir Type A, B, and C [27]. ............. 13

Fig. 1.12: Hydraulic fracture equipments are water truck (top left), fracturing sand transport truck (top right), water storage tank (bottom left) and HF process layout (bottom right) [40, 41]. ...... 17

Fig. 1.13: Fracture orientation as a function of wellbore orientation relative to in-situ stresses orientation [42].............................................................................................................................. 18

Fig. 1.14: Estimating SRA from microseismic mapping data [45]............................................... 19

Fig. 2.1: (a) Tensile failure and (b) shear failure [1]. ................................................................... 26

Fig. 2.2: (a) Triaxial strength test with β is the angle between failure plane with σ3, (b) a series of triaxial tests at different effective confining pressure (usually flattens as confining pressure increase), and linear simplification of the Mohr-Coulomb failure envelope [4, 6]. ..................... 27

Fig. 2.3: (a) Stress relationships for shear failure Mohr circle on Mohr-Coulomb failure envelope and (b) typical failure characteristics of intact rock plotted in terms of Mohr circle and Mohr-Coulomb failure envelope [5]. ...................................................................................................... 29

Fig. 2.4: Reservoir contact comparison between vertical well, unstimulated horizontal well, and multi-fractured horizontal well [17].............................................................................................. 32

Fig. 2.5: Completion types for stimulation: (a) openhole completion, (b) perforated or slotted liner, (c) blank liner with very limited clustered perforations, (d) casing packer, and (e) fully cemented liner [21]........................................................................................................................................ 36

xi

Page 13: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 2.6: (a) The PKN and (b) KGD fracture models [25]. ........................................................... 37

Fig. 2.7: Hydraulic fracture permeability measurement with laboratory experiments: (a) GRI technique [41], (b) pressure-pulse decay [42], and (c) steady-state technique [43]. .................... 43

Fig. 2.8: (a) Mechanical testing apparatus and hydraulic testing apparatus, and (b) hydraulic fracture intersecting a natural fracture [44]. ................................................................................. 45

Fig. 2.9: (a) Hydraulic fracture propagates from the tip of natural fractures and (b) hydraulic fracture propagates from weak point along natural fracture [45]. ................................................ 46

Fig. 2.10: Microseismic downhole monitoring using downhole receiver array during hydraulic fracture [47]. ................................................................................................................................. 48

Fig. 2.11: Surface microseismic monitoring during hydraulic fracture (red lines represented travel time and blue lines represented surface arrays) [52]. ................................................................... 49

Fig. 2.12: Realization of heterogeneous porous medium [68]. ..................................................... 53

Fig. 2.13: Fracture compressibility as a function of net stress on fracture [72]. .......................... 55

Fig. 2.14: Early bilinear flow within the fracture and formation [75]. ......................................... 55

Fig. 2.15: Early linear flow from the formation to the fracture [75]. ........................................... 56

Fig. 2.16: Early radial flow from the formation to the fracture [75]. ........................................... 56

Fig. 2.17: Compound linear flow from the unstimulated reservoir region to the stimulated reservoir volume [75]. .................................................................................................................................. 57

Fig. 2.18: Late radial flow around the multifractured horizontal well [75]. ................................. 57

Fig. 2.19: Trilinear model schematic in multi-fractured horizontal [78, 79]. ............................... 58

Fig. 2.20: Horizontal well multifractured enhanced fracture model schematic [80]. ................... 59

Fig. 2.21: Enhanced fracture region model for quarter of a fracture [80]. ................................... 59

Fig. 2.22: (a) Biwing fracture and (b) branched fracture [81]. ..................................................... 59

Fig. 3.1: Finite element model mesh for hydraulic fracturing (SRV) simulation………………………………………………………………………………………...78

Fig. 3.2: Bottom-hole injection pressure (field) data and predicted results from FEA. ............... 81

Fig. 3.3: Pore pressure (Pa) in different steps: (a) initial condition, (b) after in-situ stresses and boundary conditions are loaded on the domain step, (c) pump step-1 second, (d) pump step-551 s, (e) pump step-1,101 s, and (f) end of injection-2,250 s for Case 1 with k=23.4 D (deformation scale factor of 3,505.7 with final displacement of 6.853e-3 m). .................................................. 82

Fig. 3.4: Pore pressure (Pa) in different steps: (a) initial condition, (b) after in-situ stresses and boundary conditions are loaded on the domain step, (c) pump step-1 second, (d) pump step-551 s, (e) pump step-1,101 s, and (f) end of injection-2,250 s for Case 2 with k=45.85 D (deformation scale factor of 3,016.07 with final displacement of 7,957e-3 m). ................................................ 83

Fig. 3.5: Effect of Young’s modulus on bottom-hole pressure for (a) Case 1 and (b) Case 2. .... 84

Fig. 3.6: (a) Pore pressure and (b) total maximum horizontal stress as a function of distance along SRV length at various injection times for Case 1 with k=23.4D (assuming Biot’s constant=1). . 86

xii

Page 14: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 3.7: (a) Pore pressure and (b) total minimum horizontal stress as a function of distance along SRV width at various injection times for Case 1 with k=23.4D (assuming Biot’s constant=1). 88

Fig. 3.8: (a) Pore pressure and (b) total vertical stress as a function of distance along SRV height above injection ports at various injection times for Case 1 with k=23.4D (assuming Biot’s constant=1). ................................................................................................................................... 89

Fig. 3.9: (a) Pore pressure and (b) total maximum horizontal stress as a function of distance along SRV length at various injection times for Case 2 with k=45.85D (assuming Biot’s constant=1).90

Fig. 3.10: (a) Pore pressure and (b) total minimum horizontal stress as a function of distance along SRV width at various injection times for Case 2 with k=45.85D (assuming Biot’s constant=1). 91

Fig. 3.11: (a) Pore pressure and (b) total minimum horizontal stress as a function of distance along SRV height above injection ports at various injection times for Case 2 with k=45.85D (assuming Biot’s constant=1). ........................................................................................................................ 92

Fig. 3.12: Pore pressure as a function of distance using the cubic law equation for (a) Case 1 and (b) Case 2. ..................................................................................................................................... 94

Fig. 3.13: Top view of the minor fractures (natural fractures) are assumed to be inclined 30o with the major fractures (hydraulic fractures) (not to scale). ................................................................ 95

Fig. 3.14: Procedures to calculate the fracture aperture, numbers and spacing. ........................... 96

Fig. 3.15: Relationship any number of fractures, fracture aperture and fracture pressure gradient for (a) Case 1 and (b) Case 2. ....................................................................................................... 98

Fig. 4.1: (a) The site location and the maximum stress direction (45o NE) and (b) hydraulic fracture treatment location near Red Deer [39]…………………………………………………………..114

Fig. 4.2: Microseismic downhole monitoring array configuration in nearby wellbore [39]. ..... 115

Fig. 4.3: Depth distribution of microseismic events from two horizontal wellbores hydraulic fracture [39]. ............................................................................................................................... 115

Fig. 4.4: Executive summary of the two treatment wellbores, observation wellbore, producing wellbore and observed microseismic events during 12 stages of hydraulic fracture [39]. ......... 116

Fig. 4.5: Bottomhole injection pressure for horizontal wellbore A during 12 stages of hydraulic fracture [39]. ............................................................................................................................... 117

Fig. 4.6: (a) Pressure drop derived from finite element analysis in Chapter 3 and (b) Stage 7 Mohr-Coulomb failure envelope for Case 1. ........................................................................................ 126

Fig. 4.7: Stage 7 SRV length versus width for Case 1. ............................................................... 128

Fig. 4.8: Stage 7 SRV length versus height for Case 1. .............................................................. 128

Fig. 4.9: Stage 7 SRV width versus height for Case 1. ............................................................... 129

Fig. 4.10: Stage 7 SRV length versus best fit width (using microseismic data) for Case 1. ...... 129

Fig. 4.11: Stage 1 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 130

xiii

Page 15: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.12: Stage 5 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 131

Fig. 4.13: Stage 8 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 132

Fig. 4.14: Stage 9 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 133

Fig. 4.15: Stage 11 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height. .............................................................................................................................. 134

Fig. 4.16: Stage 2 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 135

Fig. 4.17: Stage 3 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 136

Fig. 4.18: Stage 4 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 137

Fig. 4.19: Stage 6 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height........................................................................................................................................... 138

Fig. 4.20: Stage 10 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height. .............................................................................................................................. 139

Fig. 4.21: Stage 12 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height. .............................................................................................................................. 140

Fig. 4.22: (a) SRV dimensions differences with microseismic events for six interpreted stages and (b) SRV best fit using microseismic event data for all stages. ................................................... 141

Fig. 4.23: (a) Pressure drop from finite element analysis in Chapter 3, (b) pressure drop derived from transient analysis equation for intact natural fractures, and (c) pressure drop derived from transient analysis equation for open natural fractures. ................................................................ 142

Fig. 4.24: Mohr-Coulomb stress failure envelope of stage 7 for Case 2. ................................... 150

Fig. 4.25: Plot of Case 1 and Case 2 effective maximum horizontal stress versus effective minimum horizontal stress for intact rock and naturally fractured reservoirs............................................. 152

xiv

Page 16: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.26: SRV dimensions of Stage 7 for Case 2 SRV dimensions: (a) SRV length versus width, (b) SRV length versus width best fit with microseismic events, (c) SRV length versus height, and (d) SRV width versus height. ...................................................................................................... 153

Fig. 5.1: Production data history………………………………………………………………...167

Fig. 5.2: Geometry with edge labels displayed. .......................................................................... 172

Fig. 5.3: SRV with triangular element mesh............................................................................... 173

Fig. 5.4: Pressure contour plot in SRV during (a) 157.5 hours, (b) 2,992.5 hours, and (c) 6,300 hours production. ........................................................................................................................ 176

Fig. 5.5: Permeability contour plot in SRV during (a) 157.5 hours, (b) 2,992.5 hours, and (c) 6,300 hours production. ........................................................................................................................ 177

Fig. 5.6: Porosity contour plot in SRV during (a) 157.5 hours, (b) 2,992.5 hours, and (c) 6,300 hours production. ........................................................................................................................ 178

Fig. 5.7: Pressure profile as a function of (a) time and (b) distance from SRV boundary to wellbore...................................................................................................................................................... 180

Fig. 5.8: Permeability profile as a function of (a) time and (b) distance from SRV boundary to wellbore....................................................................................................................................... 181

Fig. 5.9: Porosity profile as a function of (a) time and (b) distance from SRV boundary to wellbore...................................................................................................................................................... 182

Fig. 5.10: Comparison of (a) field gas flow rate and (b) bottomhole pressure with new diffusion model with porosity of 0.16 and 0.17 (no changes). ................................................................... 184

Fig. 5.11: Blasingame finite conductivity fracture with median filter. ....................................... 187

Fig. 5.12: Blasingame elliptical with median filter. .................................................................... 188

Fig. 5.13: Blasingame horizontal with median filter. ................................................................. 189

Fig. 5.14: Wattenbarger with median filter. ................................................................................ 190

Fig. 5.15: Unconventional gas module square root time plot to identify the pessimistic boundary dominated flow (green vertical line), .......................................................................................... 191

Fig. 5.16: Unconventional gas module typecurve plot to identify the pessimistic boundary dominated flow (green vertical line). .......................................................................................... 192

Fig. 5.17: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with PSS dual porosity model. ........................................................................................................................... 194

Fig. 5.18: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with slabs model...................................................................................................................................................... 195

Fig. 5.19: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with cubes model...................................................................................................................................................... 196

xv

Page 17: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.20: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with sticks model...................................................................................................................................................... 197

Fig. 5.21: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with PSS model. ............................................................................................................... 199

Fig. 5.22: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with slabs model. ............................................................................................................. 200

Fig. 5.23: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with cubes model. ............................................................................................................ 201

Fig. 5.24: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with sticks model. ............................................................................................................ 202

xvi

Page 18: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

CHAPTER 1: INTRODUCTION

1.1 Tight Gas Sands

There are several types of unconventional gas reservoirs including tight gas sand, coal bed

methane, shale gas, and natural gas hydrates. Tight gas sand definitions are defined in different

ways by different organizations. Tight gas sand reservoirs were originally defined by the U.S.

government in 1978 as formations with permeability equal or less than 0.1 mD [1]. This definition

is the most commonly accepted one by the oil and gas industry today. A second definition is the

U.S. legal definition that described tight gas sands that have an averaged un-stimulated initial gas

rate less than a maximum specified value for a given depth [2]. Tight gas sand reservoir properties

are also defined as a function of the reservoir properties including pressure, fluid properties,

reservoir and surface temperature, permeability, net pay, drainage and wellbore radius, skin and

non-Darcy constant [3]. The research documented in this thesis focuses on a study of a tight gas

sand reservoir. Given the low permeability of the reservoir and the requirement to raise its

permeability, hydraulic fracturing from horizontal wells is most used to stimulate these reservoirs.

Unconventional reservoir characteristics such as hydrocarbon amount and qualification can be

explained simply by using a resource pyramid [4]. The resource pyramid was improved by

suggesting that high-grade natural resources occupied the peak and as the grade decreased the

hydrocarbon amount increased [5]. Unconventional gas reservoir qualification also followed the

resource pyramid which meant that poorer reservoir characteristics would decrease the reservoir

permeability as shown in Figure 1.1 [2]. Low-permeability reservoirs had larger size than high-

quality reservoirs.

1

Page 19: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 1.1: Resource pyramid for natural gas [2].

Tight gas production in the U.S. has been impacted greatly by natural gas research and

successful technology application in the form of hydraulic fracturing [6]. The tight gas production

curve from the Greater Green River and the Piceance Basins, displayed in Figure 1.2, showed a

large positive increase in 1985 following about $165 million combined investment in research by

the Department of Energy (DOE) and Gas Research Institute (GRI) [6].

Fig. 1.2: Tight gas production in the United States showing natural gas response as a function of DOE and GRI investment in research and State and Federal production incentives [6].

2

Page 20: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The research investment produced 11 trillion cubic feet (tcf) of natural gas up to year of 2000

[6]. In addition to tight gas, other unconventional gas sources were also contributing significantly

to North America gas production as of 2014; at this time, as shown in Figure 1.3, unconventional

gas production was the largest source of gas [7].

The Harvard Business School and the Boston Consulting Group wrote a report on America’s

unconventional energy opportunity in 2014 which stated that the U.S. government needed to

encourage ongoing private and public sector research investment in cost-effective and low-carbon

energy technologies including potentially broader use of unconventional natural gas [8].

Fig. 1.3: Natural gas sources production by region with unconventional gas becoming the largest source of North America gas supply [7].

Canada was anticipated to have the same success level as that achieved in the U.S. for tight

gas reservoirs. It was supported by the vast volumes of gas in place for the tight gas reservoirs

estimated by different Canadian organizations [2]. Canada’s gas in place resources from both

conventional and unconventional reservoirs was estimated to be almost 4,000 tcf including a large

contribution from unconventional resources; the largest deposits are shown in Figure 1.4 [9]. The

4,000 tcf value was the sum of gas in place for tight gas (1,311 tcf), coal bed methane (801 tcf),

shale gas (1,111 tcf), and conventional gas (692 tcf). This data was also supported by the data from

National Energy Board of Canada that placed tight rock gas in place between 89 and 1,500 tcf

3

Page 21: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[10]. Figure 1.4 showed the tight gas projected in place mainly were located at the Western Canada

Sedimentary Basin [9].

Fig. 1.4: Tight gas resources distribution in Canada [9].

1.2 Basic Geology of Tight Gas Sand Reservoirs

1.2.1 Continuous Gas Accumulation

Folding, faulting, natural fractures, in-situ stresses, multi-layer systems, connectivity,

permeability barriers, net inter-bedded coals, and shales are some of the factors that must be

considered to estimate tight gas sand properties and gas in place [2]. The present definition of the

tight gas sand, on a geological basis, was defined as a basin-center or continuous gas accumulation

[11] with large dimensions, low permeability and no apparent boundaries that are in close

proximity to source rocks with very low recovery factors as shown in Figure 1.5 [12]. The

continuous gas accumulation was visualized as a collection of gas charged cells where these cells

could be productive with productivity changing from cell to cell depending on the extent of natural

fractures [2].

4

Page 22: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 1.5: Geology of tight gas reservoir as a continuous gas accumulation [12].

There were contrary points of view where tight gas sands occurred in low-permeability

reservoirs in conventional structural, stratigraphic or combination traps [13]. It was explained that

if the conventional theory was correct, then it would produce less amounts of gas than that

anticipated from the estimated volumes of gas in place [2]. The continuous gas accumulation was

first described as having characteristic pressures as either sub-normal or super-normal [14].

Pressure-depth data used to determine sub-normal and super-normal pressure profiles of the

continuous gas accumulations in North America revealed that the sub-normally pressured profile

occurred in Upper and Lower Cretaceous rocks of Alberta, Canada and Lower Silurian rocks of

Eastern Ohio, U.S. whereas super-normally pressured reservoirs occurred in Tertiary and Upper

Cretaceous rocks of Wyoming, U.S. Davis (1984) showed an example of sub-normally pressured

gas sand in Cadotte Formation, Elmsworth area, Alberta displayed in Figure 1.6. The examples

consistently showed no associated down-dip water which were opposed to the conventional

interpretation of gas and oil systems with gas and oil trapped above water.

5

Page 23: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 1.6: Pressure – depth plot, Elmsworth area, Cadotte formation as an example of sub-normally pressured formation [14].

The basin-centered gas accumulation characteristics are maybe sub-normally pressured, low

permeability (less than 0.1 mD), continuous gas saturation, and no down-dip water leg [2]. If any

of these properties were missing, the reservoir could not be defined as a continuous gas

accumulation but it was possible to have conventional structural traps within it [15]. The

continuous gas accumulation concept had been evaluated by many other researchers [16, 27, 18,

29, 20, 21, 22].

6

Page 24: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

1.2.2 Conventional Gas Accumulations

The contrary points of view from the continuous gas accumulation stated that low permeability

reservoirs from Greater Green River basin of Southwest Wyoming were not part of the continuous

gas accumulation and its productivity was dependent on the development of sweet spots [13]. The

gas fields in this basin occurred in low-permeability rocks that were trapped in conventional traps.

The data were examined from gas fields producing from Tertiary and Cretaceous reservoirs in the

Greater Green River Basin and it was concluded that these gas fields occurred in conventional

structural, stratigraphic [13]. They used 54 fields to support their conclusion that 100% of these

fields occurred in conventional traps

An improvement permeability jail model (unconventional reservoir model) was used to

describe their theory [13]. Byrnes (of the Kansas Geological Survey) was the first to propose the

permeability jail model [13]. The permeability jail model was used to describe the saturation region

with negligible water and gas effective permeability. The permeability jail model is illustrated in

Figure 1.7 and Figure 1.8 which compare two reservoirs with the same structural configuration:

Figure 1.7 shows a conventional reservoir whereas Figure 1.8 shows an unconventional reservoir.

Figure 1.7 and Figure 1.8 show relationships between capillary pressure, relative permeability, and

position within a trap. Figure 1.7 and Figure 1.8 represent a thin reservoir pinched out in a

structurally up-dip direction. Figure 1.7 shows that for a conventional reservoir, water production

extends down-dip to a free-water level (FWL), where in the middle part of reservoir both gas and

water are produced with water decreasing up-dip. In the up-dip portion of the reservoir is

characterized by water-free production of gas. Figure 1.8 shows that for an unconventional

reservoir, significant water production is limited to low structural positions near the FWL. Figure

1.8 shows that, in most cases, the effective permeability to water is sufficiently low so that there

is little to no fluid flow at or below the FWL (permeability jail). At up-dip portion of the reservoir,

water-free gas production is found. For the conventional reservoir in Figure 1.7, between 50 and

90% water saturations, there are non-zero values of relative permeability with respect to both gas

and water. The unconventional reservoir in Figure 1.8 has zero relative permeability with respect

to gas and water for 50% to 90% water saturations. This shows that there is a saturation region

within which neither gas nor water are able to flow. Figure 1.7 shows irreducible and critical water

can be of the same order of magnitude in the conventional reservoir. But in the unconventional

reservoir, the irreducible and critical water can be significantly different. It was concluded that the

7

Page 25: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

unconventional reservoirs such as those found in the Greater Green River Basin were not examples

of basin-center or continuous gas accumulations [13].

8

Page 26: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 1.7: Schematic of relative permeability curves, capillary pressures, cross section and structural map for conventional reservoir [13].

9

Page 27: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 1.8: Schematic of relative permeability curves, capillary pressures, cross section and structural map for unconventional reservoir [13].

10

Page 28: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

1.3 Naturally Fractured Reservoirs

Natural fractures are discontinuities within the rock that result from stresses that exceed the

rupture strength of the rock [25]. Naturally fractured reservoirs contain fractures – they can have

a positive or negative impact on fluid flow [26]. It was important to understand the magnitude and

direction of in-situ stresses, fractures azimuth, dip, spacing, and permeability, and matrix and

fracture water saturation in naturally fractured reservoir [27]. These data help in calculations of

gas in-place distribution between matrix and fractures. The sources of naturally fractured reservoir

information come from direct (core analysis, cutting analysis, and downhole cameras) and indirect

(drilling mud log, log analysis, well testing, and production history) sources of information [26].

1.3.1 Geological Classification

Naturally fractured reservoir classification based on the geological point of view could be

classified as tectonic (fold or fault related), regional, contractional (diagenetic), and surface related

[25, 28, 29]. Figure 1.9 shows an example of regional fractures [28].

Fig. 1.9: Example of orthogonal regional fractures in Devonian Antrim shale, Michigan Basin [28].

11

Page 29: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

1.3.1.1 Pore System Classification

Naturally fractured reservoir classification based on the pore system determines the

preliminary estimate of productive reservoir porosity classes [30]. Porosity classes are defined by

pores geometry and pore size. The geometry include intergranular, intercrystalline, vuggy, and

fracture. The pore size is classified using different techniques such as Winland [31] and Aguilera

[32] techniques. The pore size classification are megaporosity (r35>10 microns), macroporosity

(r35 between 2-10 microns), mesoporosity (r35 between 0.5-2 microns), and microporosity (r35<0.5

microns). The pore size flow capacity were: megaports were able of flowing of 10,000 bpd,

macroports of 1,000 bpd, mesoports of 100 bpd, and microports of 10 bpd [33]. Naturally fractured

reservoir pore-throat aperture estimation and its relationship with oil and gas flow rates,

permeability, and porosity are shown in Figure 1.10 [34]. Figure 1.9 follows the same format

presented by [33] using Winland’s equation.

Fig. 1.10: Chart for estimating pore-throat aperture as a function of porosities and permeability and possible ranges of oil (bpd), and gas flow rates (scfd) for different pore-throat aperture [34].

12

Page 30: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

1.3.1.2 Storage Classification

Naturally fractured reservoirs were classified based on the storativity as being Type A, B or C

as shown in Figure 1.11. Based on this classification, many reservoirs became producible because

of the presence of the natural fractures. The reservoir Type A had a large amount of hydrocarbon

stored in the matrix porosity (low permeability) and small amount of hydrocarbon stored in the

fractures (high permeability) [27]. Type B had half of the hydrocarbon stored in the matrix (low

permeability) and half stored in the fractures (high permeability) [27]. Type C had all the

hydrocarbon stored in the fractures with no contribution from the matrix [27].

Fig. 1.11: Porosity distribution in naturally fractured reservoir Type A, B, and C [27].

1.3.1.3 Matrix-Fracture Interaction

Core provided excellent source for direct information for determining the interaction between

matrix and fracture [35]. The interaction is divided into (1) no mineralization within the fractures

and hydrocarbon flows freely from the matrix to the fractures, (2) some mineralization within the

fractures with limited hydrocarbon flow, (3) complete mineralization so hydrocarbon flow is very

low, and (4) vuggy fractures where parts of the reservoir had large porosities (up to 100%) in some

intervals where hydrocarbon flow is high if the vugs are connected.

13

Page 31: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

1.4 Hydraulic Fracturing

1.4.1 Hydraulic Fracturing Basic Concepts

Hydraulic fracturing enhances the permeability and drainage area of unconventional reservoirs

enabling wells to be economically viable. Hydraulic fracturing produces fractures in the rock

formation that stimulates the flow of hydrocarbons which increases volumes that can be recovered

[34]. The fractures are created by pumping large quantities of fluids at high but controlled pressure

down the wellbore into the target rock formation [35]. The process is designed to create small

cracks within the formation and propagate the fractures to a desired distance from the wellbore by

controlling the rate, pressure, and fluid injection duration. Hydraulic fracturing fluids include water

based fluids, oil based fluids, energized fluids (inert gas of N2 or CO2), multi-phase emulsions, and

acid fluids [36]. The additives are gelling agents, cross-linkers, breakers, fluid loss additives,

bactericides, surfactants, and clay control additives [36]. Hydraulic fractures can extend up to

several hundred ft away from the wellbore. Proppant is carried into the newly formed fractures to

keep them open after the pressure drops. This allows the trapped hydrocarbon to flow through the

fractures more efficiently. Some of the hydraulic fracture fluids and proppant remain in the

reservoir rock whereas some of the hydraulic fracture fluids return to the surface with the

hydrocarbon and formation water (flow-back process).

1.4.2 Hydraulic Fracturing – Industrial Practice

Floyd Farris of Stanolind Oil and Gas Corporation (Amoco) performed a study in 1947 to

establish relationships between observed well performance and treatment pressure [37]. This study

produced a better understanding of fracture breakdown pressures during water injection. From this

work, Farris formulated the idea of hydraulically fracturing a formation to enhance production

from oil and gas wells. This study was performed in 1947 in the Hugoton gas field in Grant County

(Kansas). A total of 1,000 gallon of naphthenic acid and palm oil (napalm) thickened gasoline was

injected followed by a gel breaker to stimulate a gas producing limestone formation at a depth of

2,400 ft. The well deliverability did not change appreciably but it was the start of hydraulic

fracturing.

14

Page 32: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

In 1948, the hydraulic fracture operation patent was issued by J. B. Clark of Stanolind Oil [37].

This patent gave exclusive license to the Halliburton Oil Well Cementing Company (Howco) to

carry out the new hydraulic fracturing process. Howco performed the first two commercial

hydraulic fracturing treatments on March 17, 1949 with the hydraulic fracture treatment cost of

$1,000 in Archer County, Texas. Howco used a blend of crude and gasoline and 150 lb of sand. In

the first year, 332 wells were hydraulically fractured with average production increased of about

75%. The application of hydraulic fracturing increased greatly and it reached more than 3,000

wells a month during the mid-1950s. In 2008, more than 50,000 hydraulic fractures stages were

completed worldwide at a cost between $10,000 and $6 million per well [37]. It was a common

industry practice to have from 8 to 40 hydraulic fracture stages in a single well [37]. Hydraulic

fracturing was estimated to have increased U.S. recoverable reserves of oil by at least 30% and gas

by 90% [37].

The setting of hydraulic fracturing has changed throughout the years. The common practice

for industry nowadays is to use complex wellbores such as horizontal wells to intersect a larger

interval of hydrocarbon bearing rock. The hydraulic fracture process steps are:

1. Pad stage: injection of hydraulic fracture fluid (slickwater, fracture liquid and gas if any)

to initiate hydraulic fracture creation and propagate the created fracture.

2. Slurry stage: injection of hydraulic fracture fluid and proppant to place the proppant in the

created fracture.

3. Spacer stage: injection of hydraulic fracture fluid to make sure the proppants in place.

4. Flush stage: injection of slickwater to clean the wellbore from any hydraulic fracture fluids

and proppants.

5. Stop pumping and flow-back to the well to recover any hydraulic fracture fluids from the

wellbore while leaving the proppant in place in the reservoir.

The type of hydraulic fracturing used is dependent on a number of variables [38]:

1. In-situ stresses direction.

2. Drilled well type.

3. Target formation reservoir properties.

4. Reservoir depth, thickness, temperature, and pressure.

5. Well completion type.

15

Page 33: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

6. Hydraulic fractures stages to be completed in the wellbore.

7. Choice of fracturing fluids and materials.

8. Cost of fracturing and materials.

The hydraulic fracturing best industry practices were summarized as: (1) hydraulic fracture

fluid of 60,000 gal and proppants of 100,000 lb with the largest treatments exceeding 1 million gal

fluid and 5 million lb proppant, (2) water as a fracturing fluid with gelling agent (e.g. borate gel

breaker), (3) surfactant as emulsions minimizer, (4) formation fluid and potassium chloride as clay

impact minimizer, (5) foams and alcohol as water usage enhancement, and (6) aqueous fluids as

the base fluid in approximately 96% of all fracturing treatments [37]. Metal-based crosslinked

fluid have been used since 1970s to enhance the viscosity of gelled water based fracturing fluids

for high temperature wells. Gel and chemical stabilizers have also been used in high temperature

reservoirs.

The first proppant that had been used by the industry was river sand and construction sand that

were filtered through a window screen. There are different trends in the proppant sizes from the

beginning, but 20/40 U.S. standard mesh sand (diameter of 0.42-0.84 mm) are used in

approximately 85% of the total world hydraulic fracturing jobs. Different proppants have been

evaluated throughout the years, including plastic pellets, steel shot, Indian glass beads, aluminum

pellets, high-strength glass beads, rounded nut shells, resin-coated sands, sintered bauxite, and

fused zirconium [37]. From the total proppant market usage, 80% of the market used sand, 10%

used resin-coated sand, and 10% used ceramics [39]. The concentration of proppant (lb/gal)

remained low until 1960s, when viscous fluids such as cross-linked water-based gels were

introduced [37]. Hydraulic fracture equipment commonly includes pumps, trucks, and tanks. An

example of a hydraulic fracturing operation is shown in Figure 1.12 [40, 41].

16

Page 34: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 1.12: Hydraulic fracture equipment are water truck (top left), fracturing sand transport truck (top right), water storage tank (bottom left) and HF process layout (bottom right) [40, 41].

Horizontal well stimulation creates different types of hydraulic fractures [42]. Hydraulic

fractures can be transverse, longitudinal, and oriented fractures. The type of hydraulic fracture

depends on the horizontal wellbore direction with respect to the minimum in-situ stress as shown

in Figure 1.13 [42].

17

Page 35: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 1.133: Fracture orientation as a function of wellbore orientation relative to in-situ stresses orientation [42].

1.5 Stimulated Rock Volume Concept and Application

The stimulated reservoir volume or stimulated rock volume (SRV) is the approximated 3-D

volume of the created fracture network during hydraulic fracturing in low permeability reservoir

that can be estimated from the microseismic event cloud [45]. The SRV properties depend on the

hydraulic fracture properties such as injection fluids, proppant, pressure, rate, and injection

duration. Reservoir properties such as natural fracture network, initial pressure, in-situ stresses

magnitude and direction, thickness, elastic properties (Young’s modulus and Poisson’s ratio), and

initial permeability also affect SRV properties. The reservoir properties dictate the complexity of

the SRV. In unconventional reservoirs, a large SRV dimension is required to create maximum

fracture-surface contact area with the unconventional formations through both size and density.

In 2002, there was a first discussion of a large fracture network creation during hydraulic

fracturing in the Barnett shale and it showed the relationship between treatment properties and

18

Page 36: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

SRV size [46, 47]. Figure 1.14 shows the example of estimating the stimulated reservoir area

(SRA) from the microseismic mapping data [45].

Fig. 1.144: Estimating SRA from microseismic mapping data [45].

It was found that the effective fracture network dimensions could be smaller than the SRV

dimensions [45]. The concept of SRV dimensions, hydraulic fracture injection port spacing, and

SRV permeability as well performance driver could be used to optimize the hydraulic fracture

design. The most of the reservoir engineering models that used the SRV concept did not consider

fracture mechanics [48].

1.6 Research Questions

Production of tight gas sands is important to Canada. Shale and tight gas resource production

is increasing which is helping to counterbalance the conventional resource production decline. In

2014, shale gas was 4% of total Canadian natural gas production whereas tight gas was 47% [49].

19

Page 37: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

National Energy Board forecasts by 2035, shale and tight gas productions together will represent

90% of Canada’s natural gas production [49]. In Canada, shale and tight gas production activities

are located mainly in Western Canada Sedimentary Basin. Tight gas production requires hydraulic

fractures and horizontal well completion for optimum production. The research questions that arise

from the literature review, presented in Chapter 2, are as follows:

1. How can SRV effective permeability and in-situ stresses change during hydraulic

fracturing?

2. What is the impact of Young’s modulus on SRV effective permeability?

3. What are the characteristics of the fracture network inside SRV?

4. What are the impacts of rock mechanical properties, effective stresses, and fracture fluid

injection volume on SRV dimensions?

5. How does the SRV evolve as a function of distance and time during production?

6. Can characteristics of SRV be deduced from the gas flow rate production?

1.7 Thesis Outline

This thesis consists of three research chapters in addition to the literature review and

introduction. The first research chapter describes a new method to estimate fracture characteristic

within stimulated reservoir volume using both finite element and semi-analytical approaches. The

second research chapter uses geomechanical theory to analyze a multi-stage hydraulic fracturing

operation in a tight gas sand formation. The third research chapter uses a novel application of the

pressure diffusion equation to examine the permeability of the fracture network. The final chapter

lists conclusions and recommendations that arise from the research documented in this thesis. The

units are used in this thesis are field units and other units.

20

Page 38: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

1.8 References

[1] Kazemi, H. 1982. Low-Permeability Gas Sands. J Pet Technol 34 (10): 2229-2232. SPE 11330 PA. http://dx.doi.org/10.2118/11330-PA

[2] Aguilera, R., and Harding, T. 2007. State-of-the-Art of Tight Gas Sands Characterization and Production Technology. Presented at the Petroleum Society’s 8th Canadian International Petroleum Conference (58th Annual Technical Meeting), Calgary, Alberta, Canada, 12-14 June. http://dx.doi.org/10.2118/2007-208

[3] Holditch, S. A. 2006. Tight Gas Sands. J Pet Technol 58 (06): 86-93. SPE 103356. http://dx.doi.org/10.2118/103356-JPT

[4] Gray, J. K. 1977. Future gas reserve potential Western Canadian Sedimentary Basin: 3d Natl. Tech. Conf. Canadian Gas Assoc.

[5] Masters, J. A. 1979. Deep Basin Gas Trap, Western Canada. AAPG Bulletin 63(2):152. [6] Bureau of Economic Geology: The University of Texas. 2000. Natural Gas. In-house

presentations. http://www.beg.utexas.edu/techrvw/presentations/talks/tinker/tinker01/page05.htm (Downloaded 16 December 2015).

[7] BP World Energy Outlook Booklet 2035. 2015. http://www.bp.com/content/dam/bp/pdf/energy-economics/energy-outlook-2015/bp-world-energy-outlook_booklet_2035.pdf. (Downloaded 16 December 2015)

[8] Harvard Business School and Boston Consulting Group. 2014. http://www.hbs.edu/competitiveness/Documents/america-unconventional-energy-

opportunity.pdf. (Downloaded 16 December 2015).

[9] Heffernan, K., and Dawson, F. M. 2010. An Overview of Canada’s Natural Gas Resources. CSUG report 2010. http://www.csug.ca/images/news/2011/Natural_Gas_in_Canada_final.pdf (Downloaded 16 December 2015).

[10] National Energy Board Of Canada, Canada’s Energy Future 2013: Energy Supply and Demand Projections to 2035; copyright Her majesty the Queen in Right of Canada (49). https://www.neb-one.gc.ca/nrg/ntgrtd/ftr/2013/2013nrgftr-eng.pdf (Downloaded 16 December 2015).

[11] Schmoker, J. W. 1995. U.S. Method for assessing continuous-type (unconventional) hydrocarbon accumulations, in Gautier, D.L., Dolton, G.L., Takahashi, K.I., and Varnes, K.L., eds., 1995, 1995 National assessment of United States oil and gas resources—Results, methodology, and supporting data: U.S. Geological Survey Digital Data Series DDS–30, 1 CD–ROM.

[12] Schenk, C. J., and Pollastro, R. M. 2002. Natural Gas Production in the United States; U.S. Geological Survey Fact Sheet FS-113-01, January. http://pubs.usgs.gov/fs/fs-0113-01/ (Downloaded 16 December 2015).

21

Page 39: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[13] Shanley, K., Cluff, R. M., and Robinson, J. W. 2004. Factors Controlling Prolific Gas Production From Low-Permeability Sandstone Reservoirs: Implications for Resource Assessment, Prospect Development and Risk Analysis. AAPG Bulletin (1083-1121), August.

[14] Davis, T. B. 1984. Subsurface Pressure Profiles in Gas Saturated Basins. AAPG Memoir (38): 189-203.

[15] Shirley, K. 2004. How did the Tight Gas Get Here? Debate Taps Petroleum Systems, AAPG Explorer, April.

[16] McPeek, L. A. 1981. Eastern Green River Basin: A Developing Giant Gas Supply from Deep, Overpressured Upper Cretaceous Sandstones. AAPG Bulletin (65): 1078–1098.

[17] Law, B. E., and Dickenson, W. W. 1985. Conceptual model for origin of abnormally pressured gas accumulations in low-permeability reservoirs. AAPG Bulletin (69): 1295–1304.

[18] Spencer, C. W. 1987. Hydrocarbon Generation as a Mechanism for Overpressuring in the Rocky Mountain region. AAPG Bulletin (71): 368-388.

[19] Spencer, C. W. 1989. Review of characteristics of low permeability gas reservoirs in western United States. AAPG Bulletin (73).

[20] Law, B. E., and Spencer C. W. 1989. Geology of tight gas reservoirs in the Pinedale anticline area, Wyoming, and at the multiwell experiment site, Colorado. U.S. Geological Survey Bulletin (1886).

[21] Surdam, R. C. 1997. A New Paradigm for gas Exploration in Anomalously Pressured ‘‘Tight-Gas Sands’’ in the Rocky Mountain Laramide Basins, in R. C. Surdam, ed., Seals, traps, and the petroleum system. AAPG Memoir (67): 283–298.

[22] Selley, R. C. 1998. Elements of Petroleum Geology, second edition. Waltham: Academic Press.

[23] Law, B. E., and Curtis J. B. 2002. Introduction to unconventional petroleum systems. AAPG Bull., 86 (11):1851–1852.

[24] Zou, C. N. 2012. Unconventional Petroleum Geology. Beijing: Elsevier (373). [25] Stearns, D. W. 1994. AAPG Fractured Reservoirs School Notes 1982-1994, Great Falls,

Montana. [26] Aguilera, R. 2010. Naturally Fractured Reservoir Courses Notes, University of Calgary,

Calgary, Alberta, Canada. [27] Aguilera, R. 2003. Geologic and Engineering Aspects of Naturally Fractured Reservoirs.

CSEG Recorder, February. [28] Nelson, R. 1985. Geologic Analysis of Naturally Fractured Reservoirs, Contributions in

Petroleum Geology and Engineering, Vol. 1, Texas: Gulf Publishing Co. [29] Aguilera, R. 1998. Geologic Aspects of Naturally Fractured Reservoirs. The Leading Edge,

December: 1667-1670. [30] Coalson, E. B., Hartmann, D. J., and Thomas, J. B. 1985. Productive Characteristics of

Common Reservoir Porosity Types. Bulletin of the South Texas Geological Society 15 (6): 35-51.

[31] Kolodzie, S., Jr. 1980. Analysis of Pore Throat Size and Use of the Waxman-Smits Equation to Determine OOIP in Spindle Field. Paper SPE 9382 was presented at the Colorado Society of Petroleum Engineers 55th Annual Fall Technical Conference.

[32] Aguilera, R. 2002. Incorporating Capillary Pressure, Pore Throat Aperture Radii, Height Above Free Water Table, and Winland r35 Values on Pickett Plots. AAPG Bulletin 86 (4): 605-624.

22

Page 40: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[33] Martin, A. J., Solomon, S. T., and Hartmann, D. J. 1997. Characterization of Petrophysical Flow Units in Carbonate Reservoirs. AAPG Bulletin 83 (7): 734-759.

[34] U.S. Environmental Protection Agency. 2015. Hydraulic Fracturing Background Information. 9 May. http://water.epa.gov/type/groundwater/uic/class2/hydraulicfracturing/wells_hydrowhat.cfm

[35] Geological Society of America. 2015. GSA Critical Issue: Hydraulic Fracturing. http://www.geosociety.org/criticalissues/hydraulicFracturing/glossary.asp#Hydraulicfracturing

[36] Montgomery, C. 2013. Fracturing Fluids, Effective and Sustainable Hydraulic Fracturing. Dr. Rob Jeffrey (Ed.), ISBN: 978-953-51-1137-5, InTech, DOI: 10.5772/56192.

[37] Montgomery, C. T., and Smith, M. B. 2010. Hydraulic Fracturing-History of an Enduring Technology. Society of Petroleum Engineers, JPT Online.

[38] CSUG (Canadian Society for Unconventional Gas). 2015. http://www.csug.ca/images/CSUG_publications/CSUG_HydraulicFrac_Brochure.pdf. (Downloaded 17 December 2015).

[39] Cadre Proppant. 2012. http://www.cadreproppants.com/files/120724%20Denver%20Conference%20Cadre%20Proppants.pdf. (Downloaded 17 December 2015).

[40] Conoco Phillips Canada. 2015. http://www.conocophillips.ca/technology-and-innovation/unconventional/Pages/default.aspx. (Downloaded 17 December 2015).

[41] Driscoll, M. 2013. Proppant Prospects for Bauxite. 19th Bauxite & Alumina Seminar, Miami, USA, 13-15 March. http://www.indmin.com/Stub.aspx?StubID=4061 (Downloaded 17 December 2015).

[42] Abass, H. H., Soliman, M. Y., Tahini, A. M., Suriaatmadja, J., Meadows, D. L., and Sierra, L. 2009. Oriented Fracturing: A New Technique to Hydraulically Fracture Openhole Horizontal Well. Paper SPE 124483 was presented at the 2009 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, USA, 4-7 October.

[43] Hubbert, M. K., and Willis, D. G. 1957. Mechanics of hydraulic fracturing, Trans. Am. Inst. Min. Metall. Pet. Eng., 210: 153–168.

[44] Jaeger, J. C., Neville G. W. Cook, and Robert Wayne Zimmerman. 2007. Fundamentals of rock mechanics. Malden, MA: Blackwell Pub.

[45] Mayerhofer, M. J., Lolon, E. P., Warpinski, N .R., Cipolla, C. L., Walser, D., and Rightmire C. M. 2010. What is stimulated reservoir volume? SPE Prod & Oper 25(01): 89-98. SPE-119890. http://dx.doi.org /10.2118/119890-PA

[46] Fisher, M. K., Wright, C. A., Davidson, B. M., Goodwin, A. K., Fielder, E. O., Buckler, W. S., and Steinsberger, N. P. 2002. Integrating Fracture Mapping Technologies to Optimize Stimulations in the Barnett Shale. Paper SPE 77441 was presented at the SPE Annual Conference and Exhibition, San Antonio, Texas, 29 September–2 October.

http://dx.doi.org/10.2118/77441-MS

23

Page 41: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[47] Maxwell, S. C., Urbancic, T. I., Steinsberfer, N., and Zinno, R. 2002. Microseismic Imaging of Hydraulic Fracture Complexity in the Barnett Shale. Paper SPE 77440 was presented at the SPE Annual Conference and Exhibition, San Antonio, Texas, 29 September–2 October. http://dx.doi.org/10.2118/77440-MS

[48] Cipolla, C., and Wallace, J. 2014. Stimulated Reservoir Volume: A Misapplied Concept? Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 4-6 February.

[49] Natural Resources Canada. 2015. Exploration and Production of Shale and Tight Resources. http://www.nrcan.gc.ca/energy/sources/shale-tight-resources/17677

24

Page 42: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

CHAPTER 2: LITERATURE REVIEW

2.1 Failure Mechanics

A rock fails when large stress is applied to the rock and produces permanent change of rock

shape and its integrity. The failure state is usually accompanied with much lower capability to

carry loads [1]. The stress level when the rock fails is called the rock strength and it is usually

determined in the laboratory using uniaxial or triaxial tests or Brazilian test [2].

2.1.1 Tensile Failure

Tensile failure happens when the effective stress across a plane within the rock exceeds a

critical limit referred to as the tensile strength. The tensile strength for most rocks is low (typically

of order of a few MPa) and when there are natural fractures in the rock, the tensile strength, T0, is

expected to be close to zero [3]. The minimum effective stress, σ3’, is given by [1]:

𝜎𝜎3′ = −𝑇𝑇0 (1)

A hydraulic fracture is a form of tensile failure that occurs when the fluid pressure exceeds the

sum of the minimum total stress and the tensile strength of the rock [1]. Tensile failure extension

occurs when the injection pressure is higher than the minimum stress [4]. Continuous pumping of

fluid into the rock at high pressure causes the fracture to grow in the direction of the least resistance

which is the direction normal to the minimum stress. An example is illustrated in Figure 2.1a [1].

25

Page 43: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 2.1: (a) Tensile failure and (b) shear failure [1].

2.1.2 Shear Failure

Shear failure happens if the shear stress along some planes in the rock is high enough and it

develops a failed zone along the failure plane where the two sides of the plane move relative to

each other as shown in Figure 2.1b [1].

Shear failure can be determined by using a Mohr-Coulomb failure envelope. T The failure

envelope is built from the cohesion as the intercept and the internal friction angle as the slope [5].

High cohesion and internal friction angle are typical of strong rocks which are hard to fail. A

naturally fractured reservoir is a relatively weak rock with lower cohesion which is easier to fail

than strong rocks. The Mohr circle consists of maximum and minimum effective stresses [5].

The Mohr-Coulomb failure envelope was produced by using test results from triaxial tests

(Figure 2.2a and Figure 2.2b) [4, 6]. Triaxial tests involve applying a load on the sample (σ1) while

the confining pressure (σ3) is held constant until the sample fails. The Mohr-Coulomb failure

envelope slope usually decreases for most rocks as the confining pressure increases (Figure 2.2b).

But for most rocks, it is allowable to consider a linearized Mohr-Coulomb failure envelope as

shown in Figure 2.2b [4]. The linearized Mohr-Coulomb failure envelope criterion is given by [4]:

26

Page 44: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

𝜏𝜏 = 𝐶𝐶 + 𝜎𝜎𝑛𝑛𝜇𝜇𝑖𝑖 (2) 𝜏𝜏 = 𝐶𝐶 + 𝜎𝜎𝑛𝑛𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 (3)

(a)

(b)

Fig. 2.2: (a) Triaxial strength test with β is the angle between failure plane with σ3, (b) a series of triaxial tests at different effective confining pressure (usually flattens as confining pressure increase), and linear simplification of the Mohr-Coulomb failure envelope [4, 6].

27

Page 45: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

where τ is the shear stress, C is the rock cohesion, σn is the normal stress, µi is the slope of the

failure envelope, and ϕ is the internal friction angle. The normal stress on a failure plane is inclined

at an angle β to the least stress 𝜎𝜎3, where the minor principal stress is 𝜎𝜎3 as shown in Figure 2.2a,

is [1, 4]:

𝜎𝜎𝑛𝑛 = 𝜎𝜎1+𝜎𝜎32

+ 𝜎𝜎1−𝜎𝜎32

𝑐𝑐𝑐𝑐𝑐𝑐2𝛽𝛽 (4)

𝜏𝜏 = 𝜎𝜎1−𝜎𝜎32

𝑐𝑐𝑠𝑠𝑡𝑡2𝛽𝛽 (5)

2𝛽𝛽 = 90𝑜𝑜 + 𝑡𝑡 (6)

𝛽𝛽 = 45𝑜𝑜 + 𝜙𝜙2 (7)

The shear failure criterion is met when the Mohr circle touches the Mohr-Coulomb failure line.

This criterion shows a linear relationship between the effective stresses with the rock cohesion and

the internal friction angle as shown in Figure 2.3a. The triangle OAB in Figure 2.3a produces:

𝑐𝑐𝑐𝑐𝑐𝑐𝑡𝑡 =12

(𝜎𝜎1−𝜎𝜎3)

�𝐶𝐶+12(𝜎𝜎1+𝜎𝜎3)𝑡𝑡𝑡𝑡𝑛𝑛𝜙𝜙� (8)

28

Page 46: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 2.3: (a) Stress relationships for shear failure Mohr circle on Mohr-Coulomb failure envelope and (b) typical failure characteristics of intact rock plotted in terms of Mohr circle and Mohr-Coulomb failure envelope [5].

29

Page 47: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Rearranging Equation (8) yields:

12

(𝜎𝜎1 − 𝜎𝜎3)𝑐𝑐𝑐𝑐𝑐𝑐𝑡𝑡 = 𝐶𝐶𝑐𝑐𝑐𝑐𝑐𝑐𝑡𝑡 + 12

(𝜎𝜎1 + 𝜎𝜎3)𝑐𝑐𝑠𝑠𝑡𝑡𝑡𝑡 (9)

𝜎𝜎1 = 2𝐶𝐶𝐶𝐶𝑜𝑜𝐶𝐶𝜙𝜙(1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙) + 𝜎𝜎3(1+𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙)

(1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙) (10)

Shear failure of rocks during hydraulic fracturing is induced by the increase of pore pressure. There

is an alternative method to produce a shear failure envelope by plotting the effective maximum

horizontal stress and the effective minimum horizontal stress to support the results of shear failure

from the Mohr-Coulomb failure envelope [5].

Warpinski and Teufel (1987) found that shear failure occurred during hydraulic fracturing at

some distance from the center of the main hydraulic fracture [7]. Geomechanical analysis was used

to predict the shear failure and revealed that the extent of the shear failure zone was affected by

the fracture pressure. The results were verified by using microseismic monitoring data.

Jupe et al. (1993) explained that the microseismic data of a geothermal site in Urach, Germany

showed that the dominant mechanism that enhanced SRV permeability was shear failure along

pre-existing natural fractures [8].

Duchane (1998), by using microseismic observations and geological information of several

geothermal sites (such as in Los Alamos, New Mexico and Falkenberg, Sweden), showed that the

hydraulic vertical fractures were created from reopened sealed natural fractures (shear failures)

instead of induced new hydraulic vertical fractures (tensile fracture) [9].

Rahman et al. (2002) developed a model to integrate tensile-induced hydraulic fractures and

shear failure of natural fractures [10]. Their results revealed that the enhanced permeability of the

SRV was 30 times that of the initial formation permeability [10].

Warpinski et al. (2004) found that in a water saturated reservoir that a shear failure zone

extended further away from the main hydraulic fracture to create a wider SRV [11].

Palmer et al. (2007) discovered that the shear failures occurred at natural fractures far from the

central hydraulic fracture during hydraulic fracturing in the Barnett shale due to remote slippage

of natural fractures. They also determined that the enhanced permeability induced from hydraulic

30

Page 48: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

fracturing during production might not be equal because the shear or tensile fractures that were

induced during the operation might partially close during production [12].

2.2 Hydraulic Fracturing

2.2.1 Hydraulic Fracturing from Vertical and Horizontal Wells

Wellbore completion technology for hydraulic fracturing has been developed greatly since the

1980s. The wellbore completion type consists of vertical, horizontal, and multilateral wellbore. Elf

Aquitaine, in the early 1980s, had reported major success through horizontal wellbore on a low

permeability and naturally fractured reservoir at Prudhoe Bay and Rospo Mare field, offshore Italy

[13]. The Rospo Mare field was a good candidate for horizontal well completion because of the

oil was contained in fractures and vugs with a very low permeability formation. A horizontal well

was more appropriate to intersect many natural fractures systems in such formations.

Giger et al. [13] defined the criteria for drilling horizontal wells in preference to vertical wells

were defined as follows: (1) tight reservoirs especially if vertical natural fractures were present,

(2) thin formation, and (3) soft formations such as chalk which were liable to collapse [13].

Borisov [14] conducted field case studies to compare the productivity index between horizontal

and vertical wellbores. The results showed that the productivity improvement would rarely be

more than a factor of 5 except in the case of naturally fractured reservoirs. This study was

supported by other studies from Mukherjee and Economides [15].

Soliman et al. [16] explained that even though an unstimulated horizontal well might have been

successful in naturally fractured reservoirs and in reservoirs with water or gas coning problems,

there were conditions where fracturing a horizontal well might be a good option.

Bobrosky [17] compared reservoir contact between vertical, unstimulated horizontal, and

multi-fractured horizontal wellbores with a maximum reservoir contact was achieved by multi-

fractured horizontal wellbore. Figure 2.4 displays a comparison of the contact area achieved by

using a vertical well, an unstimulated horizontal well, and a hydraulically fractured horizontal well

[17]. The analysis reveals that the hydraulically fractured horizontal well has a contact area

roughly 1,000 times that of the vertical well.

31

Page 49: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 2.4: Reservoir contact comparison between vertical well, unstimulated horizontal well, and multi-fractured horizontal well [17].

Soliman et al. [16] mentioned that fracturing a horizontal well might dictate how the wellbore

might be completed and oriented. They summarized the situations for fracturing a horizontal well

might be a good option for these conditions: (1) restricted vertical flow caused by low vertical

permeability, (2) low formation productivity due to low formation permeability, (3) natural

fracture occurrence in a direction different from induced hydraulic fractures, and (4) low stress

contrast between pay zone and surrounding layers. The important parameters to be considered

during fracturing a horizontal and vertical wellbores were rock mechanics, reservoir engineering,

and operational aspects [16].

The stress distribution around a horizontal wellbore follows the same equations used in vertical

wells. The equations for radial, tangential, and vertical stresses for a horizontal wellbore parallel

to the minimum horizontal stress with assumption of no pore pressure penetration occurs are given

by [18]:

32

Page 50: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

For σH> σV

𝜎𝜎𝑟𝑟 = 𝑝𝑝𝑖𝑖 (11)

𝜎𝜎𝜃𝜃 = 3𝜎𝜎𝑉𝑉 − 𝜎𝜎𝐻𝐻 − 𝑝𝑝𝑖𝑖 (12)

𝜎𝜎𝑧𝑧 = 𝜎𝜎ℎ − 2𝑣𝑣(𝜎𝜎𝐻𝐻 − 𝜎𝜎𝑉𝑉) (13)

For σH<σV

𝜎𝜎𝑟𝑟 = 𝑝𝑝𝑖𝑖 (14)

𝜎𝜎𝜃𝜃 = 3𝜎𝜎𝐻𝐻 − 𝜎𝜎𝑉𝑉 − 𝑝𝑝𝑖𝑖 (15)

𝜎𝜎𝑧𝑧 = 𝜎𝜎ℎ − 2𝑣𝑣(𝜎𝜎𝑉𝑉 − 𝜎𝜎𝐻𝐻) (16)

where σr, σθ, σz are the effective normal stresses, pi is the wellbore pressure, σH, σh, σV are the in-

situ stresses, and v is the Poisson’s ratio

El Rabaa [19] explained that if a wellbore is drilled at an angle with respect to the maximum

stress, then multiple transverse fractures might be created. If the horizontal wellbore was drilled

parallel with the maximum stress, the created longitudinal fracture would propagate along the

wellbore [19]. Longitudinal fractures have not been as popular as the transverse fracture because

of the less reservoir contact that they yield.

2.2.2 Stress Interference due to Hydraulic Fracture Presence

Hydraulic fracture creation could change the stress field in its surroundings and potentially

affect subsequent new hydraulic fractures.

2.2.2.1 Stress Interference due to Semi-Infinite Fracture

Sneddon and Elliot [20] conducted research on stress interference caused by hydraulic

fracturing specifically studying the stress distribution in the neighborhood of the crack in an elastic

medium. They simplified the problem by assuming the crack was rectangular with limited height

while the crack length was infinite and the crack width was extremely small compared to its height

and length. In their analysis, the crack was open due to internal pressure. They developed equations

to predict the changes in three principal stresses. Their results suggested that if the stress contrast

between the in-situ stresses was very large, then the effect of stress interference might not affect

33

Page 51: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

the fracture initiation and direction. If the stress contrast was not large, then it was possible that

the preferred fracture direction might be different. The stress direction change was a function of

stress contrast, fracture height, and distance between created fractures. The study would be more

realistic if it was for a finite fracture.

2.2.2.2 Stress Interference due to Penny-Shaped Fracture

Soliman et al. [16] presented the simplified finite fracture by a penny-shaped fracture (circular

fracture). This case would be closely similar to the case of small multi-fractures intersecting a

horizontal wellbore (multiple transverse fractures) where the fracture height and length were

approximately equal. Sneddon [20] presented a mathematical solution for the stress distribution in

the neighborhood of a penny fracture for three dimensional case.

2.2.2.3 Stress Interference due to Multiple Fractures in Horizontal Well

In general, hydraulic fractures are created transverse to the wellbore to achieve maximum

reservoir contact. Therefore, it is important to understand stress changes during hydraulic

fracturing along a horizontal wellbore. Soliman et al. [16] stated that due to the creation of multiple

transverse propped open fractures, it was expected that the effect of stress interference grew as the

number of fractures increased. They used the penny-shaped fracture model to calculate the effect

of multiple fractures on the stress distribution [16]. The conducted study used the calculated stress

for 5 hydraulic fractures and for distance to fracture diameter ratios from 0.5 to 1 [16]. The results

showed that if the distances between the fractures were equal to the fracture diameter, then while

creating a fourth fracture, it would be expected that net pressure increased by about 21% above the

net pressure encountered during creation of the first fracture [16]. The analysis showed that the

same net pressure was expected to occur during creation of the fifth fracture [16]. If the distance

between fractures was half that of the fracture diameter, then the expected net pressure during

creation of the third fracture was about twice that encountered during creation of the first fracture

[16]. It was also found that the interference between fractures caused changes in all in-situ stresses

[16]. The minimum horizontal stress (perpendicular to the fracture) grew by a larger degree than

the other two stresses [16].

34

Page 52: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2.2.3 Hydraulic Fracturing on Horizontal Well Completion

The horizontal well completion design affects well performance. Hydraulically fractured

horizontal well completions are typical based on McDaniel and Willet’s (2002) guidelines, as

shown in Figure 2.5, given as follows [21]:

1. Open-hole completion is the cheapest and simplest technique. It is used when wellbore

instability is not a problem. The open-hole completion does not have control over fracture

placement in most cases.

2. Perforated or slotted liner is used when the wellbore instability is a concern or if the

operator requires limited control of fluid placement during completion or workover. Pre-

perforated or slotted liners offer significant improvement over open-hole completions but

may have poor distribution of fractures and they allow placement of larger stimulation fluid

volume than that of open-hole completions.

3. Blank liner with limited clustered perforations (un-cemented) is used mostly to improve

hydraulic fracturing in a horizontal well by allowing fracture fluid to reach all desired

intervals. By placing a number of perforations in the liner and plugging the toe, an operator

can choose where fluids exit the liner for better control of fracture placement. The lowest

cost application of blank liner is to use it as a retrievable treating string instead of a

permanent liner. During proppant fracturing treatments, some wellbore conditions can

increase the risk of getting the liner stuck and not being able to be retrieved.

4. Casing packers are more costly than other techniques. A casing packer completion features

inflatable packers that clamp onto the liner.

5. Partially cemented liners are used with an inflatable casing packer near the middle of the

horizontal where the un-cemented section of the liner is pre-perforated with a limited

number of clustered perforations to place well-distributed fractures.

6. Cemented casings are done by cementing the entire length of horizontal wellbore. It

provides the most control over fracture placement.

35

Page 53: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 2.5: Completion types for stimulation: (a) openhole completion, (b) perforated or slotted liner, (c) blank liner with very limited clustered perforations, (d) casing packer, and (e) fully cemented liner [21].

2.2.4 Simple Theories for Hydraulic Fracturing

To optimize the design of a hydraulic fracturing job, it is necessary to predict the growth of the

SRV versus the hydraulic fracturing injection parameters. SRV growth in heterogeneous

formations with low ratios of the stresses remains difficult to predict with a high level of certainty.

36

Page 54: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Throughout the years, various models have been developed to approximate fracture geometry.

Fracture models can be separated into a two-dimensional (2D) and a three-dimensional (3D)

categories. The KGD (Khristianovitch and Zheltov 1955, Geertsma and de Klerk 1969) [22, 23]

and PKN (Perkins and Kern 1961) [24] models are the most popular 2D models; both are depicted

in Figure 2.6 [25].

Fig. 2.6: (a) The PKN and (b) KGD fracture models [25].

Barree (2009) stated that in all 2D models only the fracture width and length were derived

from the models whereas the fracture height remains constant. The PKN and KGD models both

used the Sneddon (1946) [20] equation. Sneddon (1946) proposed an equation for the width of a

crack as follows [20]:

𝑤𝑤 = 2𝑢𝑢 = 4�1−𝑣𝑣2�𝑃𝑃0𝜋𝜋𝜋𝜋

𝑐𝑐 (17)

where w is the crack width, u is the crack half width, v is the Poisson’s ratio, P0 is the applied

pressure, E is the Young’s modulus, and c is the crack half length. Nordgren (1970), Barree (2009),

and Rahman and Rahman (2010) summarized the different assumptions for the PKN and KGD

models [25, 26, 27]. The PKN model assumptions are:

1. The crack length is larger than the crack height.

2. The crack height is restricted to a given section due to the existence of upper and lower

barriers.

37

Page 55: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3. There is no vertical extension in each vertical section; therefore the fracture shape is

elliptical.

4. A 2D plane strain deformation in the vertical plane is assumed.

The PKN equation provides an estimate of the crack width [24]:

𝑤𝑤 = 𝑢𝑢 = 2�1−𝑣𝑣2��𝑃𝑃𝑓𝑓−𝜎𝜎3�𝜋𝜋

𝐻𝐻 (18)

where H is the crack height.

The KGD model assumptions are:

1. The crack height is larger than the crack length.

2. The crack height is constant and uniform along the entire crack length; therefore, the cross

section is rectangular.

3. The crack width is constant in the vertical direction.

4. A 2D plane strain deformation in the horizontal plane is assumed.

The KGD equation produces the crack width is following [28]:

𝑤𝑤 = 𝑢𝑢 = 2�1−𝑣𝑣2��𝑃𝑃𝑓𝑓−𝜎𝜎3�𝜋𝜋

𝑥𝑥𝑓𝑓 (19)

where xf is the crack half-length.

2.2.5 Numerical Studies of Hydraulic Fracturing

Economides and Nolte [29] summarized the available models such as planar 3D, pseudo 3D,

and general 3D models for hydraulic fracturing. The planar 3D model assumes that the fracture is

planar and perpendicular to the minimum stress. This model is applicable when the surrounding

zones have stresses lower or similar to the stresses of the formation. The pseudo 3D model is

divided into lumped and cell based models. The lumped based model has two half-ellipses joined

at the center in the vertical profile. The fracture half-length and height are calculated at each time

step with the assumed shape is elliptical. The cell based model considers the fractures as a series

of connected cells. These models do not have fixed shapes but they do not consider fully coupled

fluid flow in vertical direction to fracture geometry calculation. The general 3D model does not

38

Page 56: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

have any assumption on the fracture orientation. The fracture orientation is affected by wellbore

perforations and orientations and stress orientations.

There are several analytical pseudo 3D and general 3D models proposed recently. Fisher et al.

(2002) did hydraulic fracture diagnostic projects in the Barnett naturally fractured shale reservoir

[30]. They showed that the fracture half-length was a function of injected fluid volume where the

fracture half-length stopped growing after a significant amount of injected fluid volume was

injected [30]. The SRV half-length and width were observed by using microseismic monitoring

[30].

Maxwell et al. (2002) observed the monitored microseismic events during hydraulic fracturing

in a Barnett naturally fractured shale reservoir [31]. They discovered from the microseismic

observations that the hydraulic fracture occasionally grew at an angle to the assumed fracture

direction (the maximum stress direction) and into the neighboring wells [31]. The results also

showed that the hydraulic fractures grew at an angle because they intersected the natural fracture

network [30]. It was also discovered that the hydraulic fracture grew toward neighboring wells

because of the depleted zones around the neighboring wells [31].

Xu et al. (2009) used a semi-analytical pseudo 3D geomechanical model to study the

interaction between fractures and injected fluid volume [32]. They found that the fracture network

complexity and its dimensions were affected by the ratio of stresses within the reservoir [32].

Maxwell et al. (2010) showed that there were cases with critically stressed fractures located

close to the point of hydraulic fracture deformation [33]. These critically stressed fractures could

trigger small stress changes which resulted in remote triggering of microseismic events [33]. This

explained one of the causes of microseismic measurement uncertainties [33].

Mayerhofer et al. (2010) predicted the SRV after hydraulic fracturing using microseismic

mapping beyond a horizontal well [34]. They illustrated the SRV as a summation of several

rectangles with constant width containing microseismic events [34]. These rectangles were parallel

with maximum stress [34]. They were located between the wellbore and the farthest event in both

sides of horizontal well [34]. In order to complete the 3D modeling of SRV, the SRV height was

estimated for the individual rectangle [34]. The limitation of this method was the requirement of

adequate microseismic events and it was only applied to a particular field [34].

39

Page 57: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Weng et al. (2011) developed an analytical 3D fracture network model in a naturally fractured

reservoir to determine SRV dimensions; the model was solved by using numerical simulation [35].

Their simulation results showed that stress anisotropy, natural fractures, and internal friction angle

affected complexity of fractures network [35]. Lowering the stress anisotropy changed the fracture

dimensions from a bi-wing fracture configuration to a complex fracture network [35].

Yu and Aguilera (2012) presented an analytical 3D model to determine the SRV dimensions

after hydraulic fracture operation in an unconventional gas reservoir by using microseismic events

and pressure diffusivity equation [36]. The SRV dimensions were obtained as a function of

injection pressure, minimum pressure that triggered microseismic events, microseismic event

occurrence time, and hydraulic diffusivity coefficient [36]. They determined the hydraulic

diffusivity coefficient to calibrate the model to predict SRV dimensions [36]. The hydraulic

diffusivity coefficient could be determined from a slope of a straight line plot between distances

of microseismic event distance from the wellbore versus the square root of occurrence time of the

microseismic event [36].

Nassir et al. (2012) developed a geomechanical 3D finite element model of SRV propagating

into tight formations [37]. Their results were in agreement with the shapes of SRVs obtained from

microseismic monitoring [37]. The dimensions of the SRV were found to be affected by low rock

cohesion and high initial contrast between the minimal and maximal stresses [37]. Their simulation

results suggested that a rock with a low rock cohesion (less than 1 MPa) produced a wider SRV

[36]. Therefore, the conclusions were that large and wide SRVs would only be found in formations

where the rock was weakened by natural fractures (low rock cohesion) [37].

McClure and Horne (2013) developed a computational model that coupled fluid flow, stresses

and deformation induced by fracture opening/sliding, and fracture propagation in a 2D discrete

fracture network [38]. The model was able to couple fluid flow and earthquake models [38]. The

model was used to investigate the interaction between fluid flow, permeability evolution, and

induced seismicity during hydraulic fracture injection into a single fault [38]. Using this model,

they explained the critical importance of including the change of state of stress induced by the

deformation caused by hydraulic fracturing [38]. These stresses directly impacted the mechanism

of the hydraulic fracture propagation and the resulting fracture network properties [38]. The

40

Page 58: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

limitations of the model was a 2D model and it required the paths of newly forming fractures to

be specified in advance [38].

2.3 Laboratory Studies of Hydraulic Fracturing

The design of a hydraulic fracturing job requires rock mechanical property data. Rock

mechanical properties include in-situ stress magnitude and direction, pore pressure, elastic

modulus, and strength parameters. The main data sources to provide the rock mechanical

properties are core testing and field measurements (measurement while drilling, wireline logs,

seismic data, and well tests).

Fjaer et al. (2008) explained that logs provided continuous data versus depth with limited depth

of investigations around the wellbore [1]. Logs themselves did not directly provide the required

rock mechanical properties [1]. For example, the rock strength could not be measured from

wireline logs, but rock strength might be estimated if appropriate correlations are used [1].

Cores provide direct measurement of rock strength and static elastic properties. The tested

cores in the laboratory may not be fully representative of the study formation because they may

have been disturbed during coring and any subsequent handling. The disturbance can be overcome

with proper sample preparation procedures, test procedures, and correction procedures. Some

considerations need to be applied for different types of rocks. For example, shales need both

special preparation procedures and special test procedures. The following are the details in the

laboratory testing to provide the properties of stimulated rock volume (complex fracture network).

The fracture network is created from the interaction of the induced hydraulic fractures and the

natural fractures.

2.3.1. Laboratory Studies on Fractured Tight Sand and Shale Permeability

Cui and Glover (2014) explained that the matrix permeability in unconventional reservoirs

could be measured using several techniques such as Gas Research Institute (GRI) or pressure-

decay technique, pressure-pulse decay (PPD), and steady-state techniques as shown in Figure 2.7

[39]. This technique is required to fill gas up pore space from all directions and measured an

41

Page 59: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

average permeability value in all directions for different pore-throat sizes and it did not consider

effective stresses [39]. This technique limitations were applicable on homogeneous rock and not

considering effective stresses. Cui and Glover (2014) also state that the pressure-pulse decay

technique could measure horizontal and vertical permeability separately on core plugs [39]. It was

affected by diffusion and confining stress [39].

Singhai and Gupta (2013) stated that the steady-state technique used constant injection rate

and pressure and Darcy’s law to determine the matrix permeability in fractured unconventional

reservoir [40]. But this technique was only applicable in high conductivity rocks [40].

42

Page 60: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

(c)

Fig. 2.7: Hydraulic fracture permeability measurement with laboratory experiments: (a) GRI technique [41], (b) pressure-pulse decay [42], and (c) steady-state technique [43].

43

Page 61: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2.3.2. Laboratory Studies on Hydraulic Fracture and Natural Fracture

Interaction

There were several laboratory experimental studies done in the past to evaluate the effect of

natural fractures on the propagation of the hydraulic fractures. Lamont and Jessen (1963)

conducted a series of laboratory experiments on six different types of rocks [44]. The tests were

under triaxial compression up to 1,142 psi and with angles of approach, θ, between the hydraulic

fracture and natural fracture varying from 30o to 60o as shown in Figure 2.8 [44]. The results

showed that some of the hydraulic fracture crossed the natural fracture [44]. Other results showed

that the hydraulic fracture propagated along the natural fracture and the hydraulic fracture would

exit from the natural fracture from the weakest point of the natural fracture [44]. The weakest point

was defined as the point with a high pressure within the natural fracture to overcome the local

fracture toughness [44]. This would cause the fracture break out of the natural fracture and initiate

a hydraulic fracture [44].

44

Page 62: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 2.8: (a) Mechanical testing apparatus and hydraulic testing apparatus, and (b) hydraulic fracture intersecting a natural fracture [44].

Potluri et al. (2005) performed detailed laboratory experiments on the natural fractures and the

hydraulic fractures interaction using Warpinski and Teufel’s interaction criterion (Figure 2.9) [45].

The results showed that: (1) the hydraulic fractures crossed the natural fractures when normal

stress on the natural fractures was high relative to the rock fracture toughness, (2) the hydraulic

fractures propagated within the natural fractures then broke out from the natural fractures tip (the

pressure at natural fracture tip exceeded the net pressure required to break out), and (3) the

hydraulic fracture propagated within the natural fracture then broke out from along the natural

45

Page 63: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

fractures (the pressure in the natural fracture was high enough to overcome the local fracture

toughness) [45].

(a)

(b)

Fig. 2.9: (a) Hydraulic fracture propagates from the tip of natural fractures and (b) hydraulic fracture propagates from weak point along natural fracture [45].

Zhang et al. (2013) conducted a series of experiments to measure propped induced hydraulic

fractures and propped natural fractures conductivity in Barnett shale using a modified API

conductivity cell at room temperature were performed [46]. If the proppant permeability was

known then the propped fracture aperture could be determined [46]. The results showed that the

fracture conductivity increased with proppant size and concentration [46].

46

Page 64: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2.4 Microseismic Monitoring during Hydraulic Fracturing

A microseismic event is a micro-earthquake that happens during hydraulic fracturing. The

precise location of the microseismic event is defined as the location of a new fracture or an existing

fracture when it is reopened. The time at which to the microseismic event is detected at the receiver

is the time which it takes for the P and S waves to travel the distance from the event location to

the receiver’s location. The wave velocity models for different formations are built by using a

dipole sonic log and a perforation shot arrival time. The microseismic event location is determined

by using the distance between the sensor and the microseismic event based on the P and S wave

and also using the orientation (azimuth and dip) determined from wave propagation direction [47].

Microseismic monitoring during hydraulic fracturing is a passive measurement of

microseismic events and it provides microseismic event arrival time, location, and magnitude. The

growth of the dimensions of the SRV during hydraulic fracturing is important for hydraulic

fracturing effectiveness. There are two methods of microseismic monitoring which are downhole

and surface monitoring. Historically, Bailey in his patent, explained about these two monitoring

methods procedures [48]. The general microseismic monitoring during the hydraulic fracture are:

1. Injecting high pressurized fluid down to wellbore.

2. Increasing fluid pressure with time to cause rock to fail and creating fractures.

3. Injecting fluid continuously into the fracture to cause the fracture to propagate and

eventually the injection fluid pressure decreasing.

4. Microseismic sensors are located around the hydraulic fracture treatment wellbore to

receive the seismic wave produced by the induced fracture.

5. Seismic arrival times are measured.

6. Predetermined wave velocity and seismic arrival time are used to determine the

microseismic events locations.

7. Using a polarization analysis to determine the microseismic event orientation (azimuth and

dip).

Raleigh et al. (1976) conducted the first microseismic monitoring application of hydraulic

fracturing during what was called the Rangely experiments [49]. They conducted experiments on

controlled fluid injection to detect the induced microseismicity. They used downhole and surface

47

Page 65: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

microseismic monitoring. The downhole monitoring used arrays of geophones in a nearby

observation wellbore and the surface monitoring used surface array sensors.

Downhole microseismic monitoring is the main direct observation method to monitor

hydraulic fracture dimensions at depth as shown in Figure 2.10.

Fig. 2.100: Microseismic downhole monitoring using downhole receiver array during hydraulic fracture [47].

Wright (1998) explained that tiltmeters identify changes in the sensor’s angular position [50].

Warpinski et al. (2006) stated that the sensor was very sensitive with a sensitivity equivalent to 0.2

inch movement over a 3,000 mile range [11]. The angular position provided a measure of the earth

deformation process [11]. The sensor only measured the tilt along one axis [11]. It was required to

have two orthogonal sensors to provide a full tilt measurements (magnitude and angle) [11].

Hydraulic fracturing produced tilt signatures that were inverted to define the dimensions of the

hydraulic fractures [11].

Surface microseismic monitoring is done by placing a large number of arrays on the surface.

For example, Hall and Kilpatrick (2009) used a surface arrays consist of 1,078 stations of 12

geophones spread out in a radial pattern around the hydraulic fracture well [51]. These geophones

48

Page 66: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

were buried to a depth of one foot to get maximum signals to noise ratio by reducing rainfall

interference (Figure 2.11) [52].

Fig. 2.111: Surface microseismic monitoring during hydraulic fracture (red lines represented travel time and blue lines represented surface arrays) [52].

2.5 Stimulated Rock Volume

2.5.1 Stimulated Rock Volume Permeability Prediction

Oda (1986) predicted the effective permeability tensor of a naturally fractured reservoir by

using the geometry of the fracture network [53]. Here, they used the cubic law by modeling the

reservoir as a cubic block of a rock containing the natural fractures that were arranged into smaller

cubes. The effective permeability could be determined by using the total number of cubes and

fracture intersections, the fracture intersection length and aperture, the fluid viscosity, and the

distance between the two adjacent cubes.

Rahman et al. (2002) found that the SRV permeability was a function of the stimulation

pressure, the in-situ stresses, and the fracture density [10]. They developed a model to consider

both the fracture propagation and the shear slippage of natural fractures. They found that the

average permeability of the SRV increased sharply with the increase of the stimulation pressure

beyond a threshold value. The threshold pressure was a function of the in-situ stresses and the

natural fractures properties in the reservoir. They also found that the permeability enhancement

was nearly linearly dependent on the natural fracture density. And they emphasized the need to

49

Page 67: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

characterize the fracture density as accurately as possible to obtain realistic permeability

enhancement prediction.

Ge and Ghassemi (2011) developed a procedure to determine the SRV permeability by

matching the calculated SRV dimensions with the volumes interpreted from the microseismic [54].

They developed a relationship between the net fracture pressure with the SRV permeability and

the SRV dimensions. They initially guessed a permeability value and then for the selected net

fracture pressure, they found the pore pressure and the in-situ stress distributions. They used this

pore pressure distribution with the Mohr-Coulomb failure envelope to determine the amount of

shear failure in the formation around the hydraulically induced fracture. From the failure envelope,

they constructed the extent of the shear failure around the hydraulic fracture thus defining the SRV.

They repeated the trial and error procedure to predict the SRV permeability until the SRV

dimensions matched the microseismic SRV.

Bahrami et al. (2012) derived a semi-analytical equation to model the fracture permeability as

a function of the well test permeability, the fracture aperture, and the fracture spacing [55]. A fitted

correlation was derived based on the sensitivity analysis of the well test permeability with the

fracture parameters such as the fracture permeability, aperture, spacing and compressibility.

Johri and Zoback (2013) presented a study that showed that the fracture permeability was

enhanced during the hydraulic fracturing caused by the slip of natural fractures [56].

Nassir (2013) developed a coupled reservoir-geomechanics model to determine the SRV

permeability [57]. He observed that the maximum permeability enhancement resulted at high

injection rates.

2.5.2 Hydraulic Fracture – Natural Fracture Interaction

Interactions between hydraulic fractures and natural fractures have been studied both

experimentally and numerically since the 1960s [44]. The hydraulic fracturing experiments,

conducted by Blanton (1982), in a pre-fractured material in the laboratory under tri-axial loads

revealed that the hydraulic fractures preferred to cross the pre-existing fracture only under high

differential stresses and high angles of approach [58].

50

Page 68: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Warpinski and Teufel (1987) showed that the interaction of hydraulic fractures and natural

fractures was affected by the in-situ stresses contrast, the natural fracture spacing and permeability,

and the hydraulic fracture treatment pressure [7].

Shimizu et al. (2014) investigated the influence of natural fracture permeability and approach

angle on hydraulic fracturing simulation [59]. They found that when the angle was high and the

permeability was low, the hydraulic fracture ignored the existence of the natural fractures and the

hydraulic fracture propagated straight to the direction of the maximum in-situ stress.

Pirayehgar and Dusseault (2014) used a Universal Distinct Element Code (UDEC) to analyze

fluid injection through fractures in an impermeable reservoir [60]. It was observed that branching

occurred at a short distance from the injection point and the branching was usually suppressed

under a high in-situ stress ratio. It was also found that the natural fractures were reopened mostly

parallel with the maximum in-situ stress.

2.5.3 Pressure Drop and Fracture Aperture Estimation

Barenblatt et al. (1960) modeled the pressure drop in the naturally fractured reservoir

combining a high diffusivity continuum (fracture network) and a low diffusivity continuum

(porous rock matrix) [61]. A relationship between the liquid pressure in the matrix and the natural

fractures was proposed. The fluid transfer was evaluated between the natural fractures and the

matrix in the naturally fractured reservoir.

Kim and Schechter (2009) used a discrete fracture network model, image logs and core analysis

to estimate fracture aperture [62]. The results showed that the model could estimate fracture

aperture by incorporating the outcrop maps, the image logs, the computer tomographic imaging,

and the core fracture data.

Yu and Aguilera (2012) developed an analytical model to solve a 3D pressure diffusion

equation and predict SRV dimensions [36]. The hydraulic diffusivity coefficient was determined

to calibrate the model and then the SRV dimensions were predicted. The hydraulic diffusivity

coefficient could be determined from the slope of a straight line plot between the distances of the

microseismic event distance with the wellbore versus the square root of the occurrence time of the

microseismic event. Izadi and Elsworth (2014) used a combination of cubic law and Darcy’s law

51

Page 69: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

to determine the pressure drop along the SRV [63]. This combined relationship was a function of

the fracture initial aperture, permeability, and spacing.

Estimation of fracture network characteristics such as the fracture aperture and spacing have

not been modeled adequately in the recent studies.

2.5.4 Effect of Geomechanical Properties on the SRV

Key geomechanical properties used in the analytical models (KGD and PKN) for the hydraulic

fracture are Young's modulus and the Poisson’s ratio [64]. Bratton (2011) has shown that in-situ

stresses anisotropy dictated the complexity of the fracture network where a smaller in-situ stresses

anisotropy would not exhibit a pronounced directional preference of the hydraulic fracture network

[65]. Nassir (2013) showed cohesion to be important; the effective stresses were proportional to

the cohesion of the rock [57]:

𝜎𝜎1′ = 2𝑆𝑆0𝐶𝐶𝑜𝑜𝐶𝐶𝜙𝜙1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙

+ 𝜎𝜎3′1+𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙

(27)

where 𝜎𝜎1′ is the maximum effective stress, 𝜎𝜎3′ is the minimum effective stress, S0 is the cohesion

and 𝑡𝑡 is the friction angle.

Fang et al. (2015) found a relationship between important parameters such as the bulk modulus,

critical stress intensity factor, fracture aperture, spacing, and permeability [66].

2.6 Behavior of Naturally Fractured Reservoir

Bulnes and Fitting (1945) and Imbt and Ellison (1946) had differentiated the types of porosities

in rocks [67, 68]. Void systems of sandstones were typical of primary porosity. Secondary porosity

was small in openings and it was controlled by fracturing or jointing where it was not highly

interconnected. These types of porosity could be channels or vugular voids that had been

developed during weathering or burial such as limestones or dolomites. Joints or fissures were

another types of secondary porosities in shale, siltstone, limestone or dolomite and they were

usually vertical. In most cases, the two types of porosities were found together in the rock.

52

Page 70: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The realization of this porous medium was as a complex of discrete volumetric elements with

anisotropic primary porosities coupled with secondary porosities as shown in Figure 2.12 [69].

Fig. 2.122: Realization of heterogeneous porous medium [68].

Warren and Root (1963) were the first to propose the reservoir contained the primary

(intergranular) and the secondary (fissure or vugular) porosities. They assumed the primary

porosity region contributed significantly to the pore volume but contributes insignificantly to the

flow capacity [69]. They developed an idealized model to study the behavior of dual porosity

systems (naturally fractured reservoir) with pseudo steady fluid transfer from matrix to fracture.

This study proposed two parameters to describe the deviation of the behavior of dual porosity

medium from a homogeneous porous medium. The first parameter was ω. It was a secondary fluid

capacity measurement (storativity ratio). The second parameter was λ. It was a ratio of matrix

permeability to the fracture permeability (interporosity flow coefficient). The two parameters are

as shown below:

𝜔𝜔 = 𝜙𝜙𝑓𝑓𝐶𝐶𝑓𝑓𝜙𝜙𝑓𝑓𝐶𝐶𝑓𝑓+𝜙𝜙𝑚𝑚𝐶𝐶𝑚𝑚

(28)

𝜆𝜆 = 𝛼𝛼𝑟𝑟𝑤𝑤2𝑘𝑘𝑚𝑚𝑘𝑘𝑓𝑓

(29)

where 𝑡𝑡𝑓𝑓 and 𝑡𝑡𝑚𝑚are the fracture and matrix porosity, 𝑐𝑐𝑓𝑓 and 𝑐𝑐𝑚𝑚 are the fracture and matrix

compressibility, α is the shape factor, rw is the wellbore radius, kf and km are the fracture and matrix

permeability.

53

Page 71: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Stearns (1982) defined the natural fracture as a macroscopic planar discontinuity that results

from stresses that exceeded the rupture strength of the rock [70]. Nelson (1985) defined the natural

fracture as a naturally occurring macroscopic planar discontinuity in rock due to deformation and

it could have positive or negative effects on the fluid flow [71]. Aguilera (1998) explained all

reservoirs contained at least some natural fractures but if the natural fractures effect were negligible

then the reservoir could be classified as a conventional reservoir [72]. The natural fractures were

the main production factor in a wide range of unconventional reservoirs including tight gas

reservoir.

Aguilera (2003) also mentioned that it was important to know the magnitude and orientation

of in-situ stresses; spacing, aperture, permeability and porosity of the fractures; also permeability

and porosity of the matrix [73]. This information would lead to estimates of hydrocarbon in place

and distribution between the fracture and the matrix based on the flow capacity of the fracture and

the matrix. Stearns (1982), Nelson (1985), and Aguilera (1998) classified the natural fractures from

the geological point of view as tectonic (fold or fault related), regional, contractional (diagenetic)

and surface related [70, 71, 72].

An important property of naturally fractured reservoirs was the fracture compressibility.

Aguilera (2003) stated that the fracture compressibility for zero mineralization within the fracture,

should be higher than the matrix compressibility because of the unrestricted fluid flow [73]. The

differences between these values depended on the amount of the secondary mineralization within

the fractures, the fracture orientation, the in-situ stresses and the reservoir pressure condition [73].

Aguilera’s (1998) correlation for the fracture compressibility is shown in Figure 2.13 [72].

54

Page 72: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 2.133: Fracture compressibility as a function of net stress on fracture [72].

2.6.1 Flow Regimes for Multi-fractured Horizontal Well in a Naturally

Fractured Reservoir

Chen and Raghavan (1997) proposed the flow regimes for a multifractured horizontal well in

a rectangular drainage region for two fractures [74]. They neglected wellbore storage effects. The

first flow regime was bilinear or linear flow. Bilinear flow occurred when the fracture conductivity

was finite and the fracture length was greater than the fracture height. Linear flow within the

fracture toward the horizontal well and within the formation is shown in Figure 2.14 [75].

Fig. 2.14: Early bilinear flow within the fracture and formation [75].

Nobakht et al. (2011) proposed the second flow regime to be early linear flow. It occurred

when there was linear flow from the formation toward the fractures and the flow within the

55

Page 73: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

fractures was negligible [76]. In multifractured horizontal well in unconventional reservoir, the

early linear flow was expected to be dominant and could last for years depended on the formation

permeability as shown in Figure 2.15.

Fig. 2.15: Early linear flow from the formation to the fracture [75].

The third flow regime was early radial flow is shown in Figure 2.16. This flow regime was

fluid flow from the fracture tip toward the horizontal wellbore. This flow regime depended on the

fracture length and spacing. It happened after the early linear flow and before the fracture

interference. It was only observed when the fracture was very short or far apart [74]. Early time

flow such as early radial flow within the fractures could occur when the horizontal wellbore length

was larger than the formation thickness within a range of LD ≤ 20. The fluid flowed radially from

all directions that were perpendicular to the horizontal wellbore [77].

𝐿𝐿𝐷𝐷=𝐿𝐿𝑤𝑤ℎ �

𝑘𝑘𝑧𝑧𝑘𝑘𝑥𝑥

(41)

where Lw is the horizontal wellbore length, h is the formation thickness, kz and kx are the

permeability in the z and x axis.

Fig. 2.16: Early radial flow from the formation to the fracture [75].

56

Page 74: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The fourth flow regime was the compound linear flow as shown in Figure 2.17. It occurred

once the fractures had interfered each other. The fluid flowed from the unstimulated reservoir

volume toward the stimulated reservoir volume.

Fig. 2.17: Compound linear flow from the unstimulated reservoir region to the stimulated reservoir volume [75].

The fifth flow regime was late radial flow as shown in Figure 2.18. The flow occurred around

the multifractured horizontal well boundaries. The flow pattern was similar to the late time

production of the vertically fractured well. It only occurred if the well existed all alone in an

undeveloped field and usually it required very long production time to be developed in tight

unconventional reservoirs [74, 75]. Lastly was the boundary dominated flow, it could be a pseudo

steady state flow (no flow boundaries) or steady state flow (constant pressure boundaries).

Fig. 2.18: Late radial flow around the multifractured horizontal well [75].

57

Page 75: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2.6.2 Flow Regions for Multi-fractured Horizontal Well in a Naturally

Fractured Reservoir

This section presents literature that discusses the flow regions that occurring in multifractured

horizontal wells. Ozkan et al. (2009) and Brown et al (2009) proposed a trilinear flow model where

the drainage volume of multifractured horizontal well was limited to the inner reservoirs between

the fractures (Figure 2.19) [78, 79]. The basis of the trilinear flow model was the production life

of the multifractured horizontal well that was dominated by the linear flow regimes. The trilinear

flow model coupled the linear flow in three adjacent flow regions. The flow regions were the outer

reservoir, the inner reservoir between fractures and the hydraulic fractures. They assumed uniform

distribution of identical hydraulic fractures along the length of the horizontal well.

Fig. 2.19: Trilinear model schematic in multi-fractured horizontal [78, 79].

To allow production from the inner reservoir region between the hydraulic fractures, the region

was assumed to have natural fractures (dual porosity model). The flow regime in the inner reservoir

region was assumed to be transient flow and the model used the transient interporosity coefficient.

The latest model was a horizontal well multi-fractured enhanced fracture region model from

Stalgorova and Mattar (2012) as shown in Figure 2.20 and Figure 2.21 [80]. The model used the

same concept from [78] but it assumed the unstimulated reservoir region beyond the fracture tip

(the shaded area) contribution was negligible and the unstimulated reservoir region (darker color

area) between the fractures contribution was taken into account. Stalgorova and Mattar (2012)

adapted the branch fracture concept from Daneshy (2003) [80] (Figure 2.22) [80]. It was explained

that the branched fracturing could be caused by wellbore orientation respect to in-situ stresses,

58

Page 76: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

perforation pattern and natural fractures [81]. It could also be caused by low anisotropy of in-situ

stresses.

Fig. 2.20: Horizontal well multifractured enhanced fracture model schematic [80].

Fig. 2.21: Enhanced fracture region model for quarter of a fracture [80].

Fig. 2.22: (a) Biwing fracture and (b) branched fracture [81].

59

Page 77: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2.6.3 Rate Transient Analysis (RTA) in a Naturally Fractured Reservoir There are different methods available to analyze production data. The two distinct methods are

typecurve and non-typecurve methods. Arps (1945) was the first to develop production data

analysis methods [82]. He developed decline curves for oil and gas production during transient

flow. The traditional decline analysis had limitations: it was not able to disassociate the production

forecast from operating conditions [82]. He assumed the historical operating condition stayed

constant for future production.

Fetkovich (1980) then extended the decline curve into the typecurve concept for production

data analysis, where before the typecurve concept was only used for pressure transient analysis

[83]. It was found that late time (boundary dominated flow) data could be matched to typecurves.

Both of these traditional decline curves relied on matching the model with the production data.

The limitations was the assumption that the productions parameters would remain constant with

time.

Recent methods considered variable production parameters. These included Wattenbarger

(1998), Blasingame et al. (1991), and Agarwal et al. (1998) [84, 85, 86]. The improvement on

traditional analysis was the use of a normalized rate and a pseudotime. The normalized rate used

the pressure drop (q/Δp) that allowed the effect of pressure changes to be taken into account in the

analysis. The pseudotime was the time function for gas reservoirs that took into account

compressibility changes of gas with pressure that would allow the gas material balance to be dealt

carefully as the reservoir pressure decreased with time.

Blasingame et al. (1991) provided the typecurves for radial flow, elliptical well, fractured

vertical well, horizontal well with no fractures, finite conductivity fractures and infinite

conductivity fractures [84]. The same procedure from Wattenbarger et al. (1998) was used by

plotting the logarithm of the logarithm (log log) of the normalized rate with material balance

pseudotime with options of having rate integral and rate derivative on the y axis [84, 85]. Their

limitations were the rate integral was very sensitive with early time errors and did not distinguish

the different flow regimes.

Wattenbarger et al. (1998) typecurves were used to analyze linear flow specifically in tight

reservoirs where the linear flow could be dominant and last for years [85]. It assumed a vertical

well with fractures in the center of a rectangular reservoir where the fractures were assumed to

60

Page 78: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

reach the reservoir boundaries. The log log plot of the normalized rate with material balance

pseudo time was used with another option of having pressure derivative on the y axis. Their

limitation was the fact that the typecurve was only applicable for the linear flow and not for the

boundary dominated flow. Agarwal et al. (1998) provided the typecurves for radial flow and a

fractured vertical well [86]. It used the same procedure from Wattenbarger et al. (1998) and

Blasingame et al. (1991) but with different transient characterization using dimensionless reservoir

boundaries parameters [86]. It was found that this typecurve was more unique than Blasingame et

al. (1991). But all these available typecurves are only applicable for vertical wells, both fractured

and un-fractured.

2.7 What is Missing in the Literature?

In the previous studies summarized in the literature review, investigations of some important

SRV parameter were missing. The missing SRV parameter studies are: (a) the SRV effective

permeability, in-situ stresses, and pressure drop during hydraulic fracturing and production at

reservoir conditions as a function of distance and time, (b) the impact of modulus properties on

SRV effective permeability, (c) the effect of fracturing operation parameters on the induced

hydraulic fractures and re-opening natural fractures aperture, numbers, and spacing, (d) the impact

of rock mechanical properties, effective stresses, and fracture fluid injection volume on SRV

dimensions during hydraulic fracturing, and (e) the application of nonlinear diffusivity equation

solution to solve flow rate, permeability, pore pressure, and porosity within SRV as a function of

distance and time during production.

61

Page 79: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2.8 References

[1] Fjaer, E., Holt, R. M., Horshud, P., Raaen, A. M., and Risnes, R. 2008. Petroleum Related Rock Mechanics, second edition. Amsterdam: Elsevier B. V.

[2] Claesson, J., and Bohloli, B. 2002. Brazilian Test: Stress Field and Tensile Strength of Anisotropic Rocks Using an Analytical Solution. Int. J. of Rock Mechanics & Mining Sciences 39 (2002): 991-1004.

[3] Lockner, D. A. 1995. Rock Failure. In Rock Physics and Phase Relations, first edition, T.J. Ahrens, Chap. 8, 127-147. Washington, D. C.: American Geophysical Union.

[4] Zoback, M. D. 2007. Reservoir Geomechanics, first edition. Cambridge: Cambridge University Press.

[5] Jaeger, J. C., and Cook, N. G. W. 1979. Fundamentals of Rock Mechanics, third Edition. London: Chapman and Hall.

[6] Das, M. B. 2006. Principles of Geotechnical Engineering, fifth edition. Ontario: Thomson Canada Limited.

[7] Warpinski, N. R. and Teufel, L. W. 1987. Influence of Geologic Discontinuities on Hydraulic Fracture Propagation. J Pet Technol 39 (02).

http://dx.doi.org/10.2118/13224-PA [8] Jupe, A. J., Willis-Richards, J., and Nicollis, J. D. M. 1993. Review of HDR projects. CSM

Associates Ltd., Report 1R03/06 for AEA Technology Ltd. Harwell, U.K. [9] Duchane, D. 1998. The History of HDR Research and Development. Proc., the Fourth

International HDR Forum, Strasbourg, France, 28-30 September, LA-UR--98-4067; CONF-9809136. http://www.osti.gov/geothermal/biblio/350838

[10] Rahman, M. K., Hossain, M. M., and Rahman, S. S. 2002. A Shear-Dilation Based Model for Evaluation of Hydraulically Stimulated Naturally Fractured Reservoirs. Int. J. Num. Analyt. Methods Geomech. 26:469-497.

http://dx.doi.org/10.1002/nag.208 [11] Warpinski, N. R., Wolhart, S. L., and Wright, C. A. 2004. Analysis and Prediction of

Microseismicity Induced by Hydraulic Fracture. SPE J. 9 (1): 24-33. SPE-87673-PA. http://dx.doi.org/10.2118/87673-PA

[12] Palmer, I., Moschovidis, Z., and Cameron, J. 2007. Modeling Shear Failure and Stimulation of the Barnett Shale After Hydraulic Fracture. Paper SPE 106113 was presented at the 2007 Hydraulic Fracture Technology Conference, Texas, USA, 29-31 January. SPE-106113. http://dx.doi.org/10.2118/106113-MS

[13] Giger, F. M., Reiss, L. H., and Jourdan, A. P. 1984. The Reservoir Engineering Aspects of Horizontal Drilling. Paper SPE 13024 was presented at the SPE Annual Technical Conference and Exhibition, Houston, Texas, 16-19 September. SPE 13024 MS. http://dx.doi.org.ezproxy.lib.ucalgary.ca/10.2118/13024-MS

[14] Borisov, Ju. P. 1964. Exploitation of Oil Deposits by Horizontal Wells at a Multiple Deviations. Nedra, Moscow, Institute Francais du Petrole, Rueil 28363.

[15] Mukherjee, H., and Economides, M. 1991. A Parametric Comparison of Horizontal and Vertical Well Performance. SPE Formation Evaluation 6 (02), June. SPE 18303 PA. http://dx.doi.org/10.2118/18303-PA

[16] Soliman, M. Y., East, L., and Adams, D. 2004. Geomechanics Aspects of Multiple Fracturing of Horizontal and Vertical Wells. Paper SPE 86992 was presented at the SPE International

62

Page 80: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Thermal Operations and Heavy Oil Symposium and Western Regional Meeting, Bakersfield, California, U.S.A, 16-18 March. SPE 86992. http://dx.doi.org/10.2118/86992-MS

[17] Bobrosky, D. 2010. Multi-Stage Fracturing of Horizontal Wells. Oral presentation to Peters and Co. Limited, Lake Louise, Alberta, Canada, 28-29 January. https://www.petersco.com/pdf/Multi-Stage%20Fracturing%20of%20Horizontal%20Wells%20-%20Panel%204%20-%20Packers%20Plus%20Energy%20Services.pdf (Downloaded 21 December 2015).

[18] Soliman, M. Y. 1990. Interpretation of Pressure Behavior of Fractured, Deviated, and Horizontal Wells. Paper SPE 21062 was presented at the SPE Latin American Petroleum Engineering Conference, Rio de Janeiro, 14-19 October. SPE 21062. http://dx.doi.org.ezproxy.lib.ucalgary.ca/10.2118/21062-MS

[19] El Rabaa, W. 1989. Experimental Study of Hydraulic Fracture Geometry Initiated from Horizontal Wells. Paper SPE 19720 was presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, U.S.A, 8-11 October. SPE 19720. http://dx.doi.org.ezproxy.lib.ucalgary.ca/10.2118/19720-MS

[20] Sneddon, I. N. 1946. The Distribution of Stress in the Neighborhood of a Crack in an Elastic Solid. Proc. The Royal Society of London. Series A: Mathematical and Physical Sciences 187 (1009): 229-260.

http://dx.doi.org/10.1098/rspa.1946.0077 [21] McDaniel, B. W. and Willet, R. M. 2002. Stimulation Techniques for Low-Permeability

Reservoirs with Horizontal Completions that Do Not Have Cemented Casing. Presented at the SPE Gas Technology Symposium, Calgary, Alberta, Canada, 30 April-2 May. SPE 75688. http://dx.doi.org.ezproxy.lib.ucalgary.ca/10.2118/75688-MS

[22] Khristianovitch, S. A. and Zheltov, Y. P. 1955. Formation of Vertical Fractures by Means of Highly Viscous Fluids. Presented at the 4th World Petroleum Congress, Rome, Italy, 6-15 June. WPC-6132.

[23] Geertsma, J., and de Klerk, E. 1969. A Rapid Method of Predicting Width and Extent of Hydraulically Induced Fractures. J. Pet. Tech. 21: 1571-1581.

[24] Perkins, T. K. and Kern, L. R. 1961. Widths of Hydraulic Fractures. J Pet Tech 13 (9): 937-949. SPE-89-PA.

http://dx.doi.org/10.2118/89-PA [25] Rahman, M. M. and Rahman, M. K. 2010. A Review of Hydraulic Fracture Models and

Development of an Improved Pseudo-3D Model for Stimulating Tight Oil/Gas Sand. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects 32 (15): 1416-1436.

http://dx.doi.org/10.1080/15567030903060523 [26] Nordgren, R. P. 1970. Propagation of a Vertical Hydraulic Fracture. SPE J 12 (04): 306-

314. http://dx.doi.org/10.2118/3009-PA

[27] Barree, R. D. 2009. Modeling Fracture Geometry. St. Petersburg State Polytechnical University, Institute of Applied Mathematics and Mechanics. http://mech.spbstu.ru/images/4/4d/Modeling_Fracture_Geometry.pdf (Downloaded 24 March 2015).

63

Page 81: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[28] Daneshy, A. A. Jr. 1971. Opening of a Pressurized Fracture in an Elastic Medium. Presented at the 22nd Annual Technical Meeting of the Petroleum Society of CIM, Banff, Canada, 2-4 June. PETSOC-7616.

http://dx.doi.org/10.2118/7616 [29] Economides, M. J. and Nolte, K. G. 2000. Reservoir Stimulation, third edition. West

Sussex: John Wiley & Sons Ltd. [30] Fisher, M. K., Wright, C. A., Davidson, B. M., Goodwin, A. K., Fielder, E. O., Buckler, W.

S., and Steinsberger, N. P. 2002. Integrating Fracture Mapping Technologies to Optimize Stimulations in the Barnett Shale. Paper SPE 77441 was presented at the SPE Annual Conference and Exhibition, San Antonio, Texas, 29 September–2 October. SPE-77441.

http://dx.doi.org/10.2118/77441-MS [31] Maxwell, S. C., Urbancic, T. I., Steinsberfer, N., and Zinno, R. 2002. Microseismic Imaging

of Hydraulic Fracture Complexity in the Barnett Shale. Paper SPE 77440 was presented at the SPE Annual Conference and Exhibition, San Antonio, Texas, 29 September–2 October. SPE-77440. http://dx.doi.org/10.2118/77440-MS

[32] Xu, W., Calvez, J. L., and Thiercelin, M. 2009. Characterization of Hydraulically-Induced Fracture Network Using Treatment and Microseismic Data in a Tight-Gas Formation: A Geomechnical Approach. Paper S{E 125237 was presented at the 2009 SPE Tight Gas Completions Conference, San Antonio, 15-17 June. SPE-125237.

http://dx.doi.org/10.2118/125237-MS [33] Maxwell, S. C., Underhill, B., Bennet, L., and Martinez, A. 2010. Key Criteria for a

Successful Microseismic Project. Paper SPE 134695 was presented at the SPE Annual Technical Conference and Exhibition. Florence, Italy, 19–22 September. SPE-134695. http://dx.doi.org/10.2118/134695-MS

[34] Mayerhofer, M. J., Lolon, E. P., Warpinski, N. R., Cipolla, C. L., Walser, D., and Rightmire C. M. 2010. What is stimulated reservoir volume? SPE Prod & Oper 25(01): 89-98. SPE-119890. http://dx.doi.org /10.2118/119890-PA

[35] Weng, X., Kresse, O., Cohen, C., Wu, R., and Gu, H. 2011. Modeling of Hydraulic-Fracture-Network Propagation in a Naturally Fractured Formation. SPE Prod & Oper 26(04): 368-380. SPE-140253.

http://dx.doi.org/10.2118/140253-PA [36] Yu, G., and Aguilera, R. 2012. 3D Analytical Modeling of Hydraulic Fracture Stimulated

Reservoir Volume. Paper SPE 153486 was presented at the SPE Latin American and Caribbean Petroleum Engineering Conference, Mexico City, Mexico, 16-18 April. SPE-153486. http://dx.doi.org/10.2118/153486-MS

[37] Nassir, M., Settari, A. and Wan, R. 2012. Prediction of SRV and Optimization of Fracture in Tight Gas and Shale Using a Fully Elasto-Plastic Coupled Geomechanical Model. Presented at the SPE Hydraulic Fracture Technology Conference, The Woodlands, Texas, 4-6 February. SPE-163814

http://dx.doi.org/10.2118/163814-PA

64

Page 82: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[38] McClure, M. W. and Horne, R. N. 2013. Discrete Fracture Network Modeling of Hydraulic Stimulation Coupling Flow and Geomechanics, first edition. Heidelberg, Germany: Springer International Publishing.

[39] Cui, A. X., and Glover, K. 2014. A Practical Recommendation for Permeability Measurement in Tight Sand and Shale Reservoirs. Presentation at 2014 AAPG Annual Convention and Exhibition, Houston, Texas, 6-9 April.

[40] Singhai, B. B. S., and Gupta, R. P. 2013. Applied Hydrogeology of Fractured Rocks. India: Springer-Science Business Media.

[41] Tinni, A., Fathi, E., Agarwal, R., Sondergeld, C., Akkutlu, Y., and Rai, C. 2012. Shale Permeability Measurements on Plugs and Crushed Samples. Paper SPE 162235 was presented at the SPE Canadian Unconventional Resources Conference, Calgary, Alberta, Canada, 30 October – I November.

[42] Finsterle, S., and Persoff, P. 1997. Determining Permeability of Tight Rock Samples Using Inverse Modeling. Water Resources Research 33(8): 1803-1811.

[43] PermInc. 2016. Laboratory Measurements of Relative Permeability Steady State Technique. http://perminc.com/resources/fundamentals-of-fluid-flow-in-porous-media/chapter-2-the-porous-medium/relative-permeability/laboratory-measurements-relative-permeability/steady-state-techniques/ (Downloaded 24 March 2015).

[44] Lamont, N., and Jessen, F. W. 1963. The Effects of Existing Fractures in Rocks on the Extension of Hydraulic Fractures. Paper SPE 419 was presented at the 37th Annual Fall Meeting of SPE, Los Angeles, California, 7-10 October.

[45] Potluri, N., Zhu, D., and Hill, A. D. 2005. Effect of Natural Fractures on Hydraulic Fracture Propagation. Paper SPE 94568 was presented at the SPE European Formation Damage Conference, Scheveningen, The Netherland, 25-27 May.

[46] Zhang, J., Kamenov, A., Zhu, D., Hill, A. D. 2013. Laboratory Measurements of Hydraulic Fracture Conductivities in the Barnett Shale. Paper SPE 163839 was presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 4-6 February.

[47] ESG Solutions, 2015, Microseismic Event Location. https://www.esgsolutions.com/technical-resources/microseismic-knowledgebase/microseismic-event-location (Downloaded 20 May 2015).

[48] Bailey, J. R. 1973. Mapping of Earth Fractures Induced by Hydrofracture: U. S. Patent 3 739 871.

[49] Raleigh, C. B., Healy, J. H., and Bredehoeff, J. D. 1976. An Experiment in Earthquake Control at Rangely, Colorado. Science 191 (4233): 1230-1237.

http://dx.doi.org/10.1126/science.191.4233.1230 [50] Wright, C. A. 1998. Tiltmeter Fracture Mapping From the Surface and Now Downhole. Petr

Eng Int 71 (1): 50-54. http://ezproxy.lib.ucalgary.ca/login?url=http://search.proquest.com/docview/224546974?accountid=9838

[51] Hall, M., and Kilpatrick, J. E. 2009. Surface Microseismic Monitoring of Slick-Water and Nitrogen Fracture Simulations, Arkoma Basin, Oklahoma. Presented at the SEG Houston 2009 International Exposition and Annual Meeting. SPE-132371-MS. http://dx.doi.org/10.2118/132371-MS

65

Page 83: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[52] Barker, W. 2009. Increased Production Through Microseismic Monitoring of Hydraulic Fracture Over a Multiwell Program. Presented at the 2009 SPE Annual Conference and Exhbition, New Orleans, Lousiana, 4-7 October. Paper SPE 124877.

http://dx.doi.org/10.2118/124877-MS [53] Oda, M. 1986. An Equivalent Continuum Model for Coupled Stress and Fluid Flow Analysis

in Jointed Rock Masses. Water Resource Research 22 (13): 1845-1856. http://dx.doi.org/10.1029/WR022i013p01845

[54] Ge, J. and Ghassemi, A. 2011. Permeability Enhancement in Shale Gas Reservoirs after Stimulation by Hydraulic Fracturing. Presented at the 45th US Rock Mechanics/Geomechanics Symposium, San Francisco, CA, 26-29 June. ARMA 11-514.

[55] Bahrami, H., Rezae, R., and Hossain, M. 2012, Characterizing Natural Fractures Productivity in Tight Gas. J Pet Explor Prod Technol 2 (26):107-115.

http://dx.doi.org/10.1007/s13202-012-0026-x [56] Johri, M., and Zoback, M. D. 2013. The Evolution of Stimulated rock volume during

Hydraulic Stimulation of Shale Gas Formations. Presented at the Unconventional Resources Technology Conference, Denver, Colorado, 12-14 August. URTeC 1575434.

[57] Nassir, M. 2013. Geomechanical Coupled Modeling of Shear Fracturing in Non-Conventional Reservoirs. Ph. D. Thesis, University of Calgary, Calgary, Alberta (January 2013).

[58] Blanton, T. L., 1982, An Experimental Study of Interaction between Hydraulically Induced and Pre-existing Fractures. Presented at the SPE/DOE Unconventional Gas Recovery Symposium of the Society of Petroleum Engineers, Pittsburgh, PA, 16-18 May. SPE/DOE-10847.

[59] Shimizu, H., Hiyama, M., and Ito, T. 2014. Flow-coupled DEM Simulation for Hydraulic Fracturing in Pre-Fractured Rock. Presented at the 48th US Rock Mechanics/Geomechanics Symposium, Minneapolis, 1-4 June. ARMA 14-7365.

[60] Pirayehgar, A. and Dusseault, M. B. 2014. The Stress Ratio Effect on Hydraulic Fracturing in The Presence of Natural Fractures. Presented at the 48th US Rock Mechanics/Geomechanics Symposium, Minneapolis, 1-4 June. ARMA 14-7138.

[61] Barenblatt, G. I., Zheltov, I. P., and Kochina, I. N. 1960. Basic Concepts in the Theory of Seepage of Homogeneous Liquids in Fissured Rocks (Strata). J Appl Math Mech 24 (5): 1286-1303.

http://dx.doi.org/10.1016/0021-8928(60)90107-6 [62] Kim, T. H. and Schechter, D. S. 2009. Estimation of Fracture Porosity of Naturally fractured

Reservoirs with No Matrix Porosity Using Fractal Discrete Fracture Networks. Presented at the 2007 SPE Annual Technical Conference and Exhibition, Anaheim, California, 11-14 November. SPE-110720.

[63] Izadi, G. and Elsworth, D. 2014. Reservoir stimulation and induced seismicity: Roles of fluid pressure and thermal transients on reactivated fractured networks. Geothermic 51: 368-379.

http://dx.doi.org/10.1016/j.geothermics.2014.01.014 [64] Valko, P. and Economides, M. J. 2001. Hydraulic Fracture Mechanics. West Sussex: John

Wiley and Sons.

66

Page 84: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[65] Bratton, T. 2011. Hydraulic Fracture Complexity and Containment in Unconventional Reservoirs. Oral presentation given at the ARMA Workshop on Rock Mechanics/Geomechanics, San Francisco, 23 June.

[66] Fang, Y., Elsworth, D., and Cladouhos, T. T. 2015. Estimating In-Situ Permeability of Stimulated EGS Reservoirs using MEQ Moment Magnitude: an Analysis of Newberry MEQ Data. Presented at the Fortieth Workshop on Geothermal Reservoir Engineering Stanford University, Stanford, California, 26-28 January.

[67] Bulnes, A. C., and Fitting, R. U. 1945. An Introductory Discussion of the Reservoir Performance of Limestone Formations. Trans., AIME. Vol. 160, 179.

[68] Imbt, W. C., and Ellison, S. P. 1946. API Drill and Prod. Prac. 364. [69] Warren, J. E., and Root, P. J, 1963. The Behavior of Naturally Fractured Reservoirs. SPEJ

426. [70] Stearns, D. W. 1982-1994. AAPG Fractured Reservoirs School Notes. Great Falls, Montana. [71] Nelson, R. 1985. Geologic Analysis of Naturally Fractured Reservoirs. Contributions in

Petroleum geology and engineering, Vol. 1, Gulf Publishing Co., Houston, Texas. [72] Aguilera, R. 1998. Geologic Aspects of Naturally Fractured reservoirs. The Leading Edge,

pp. 1667-1670, December. [73] Aguilera, R. 2003. Geologic and Engineering Aspects of Naturally Fractured Reservoirs.

CSEG Recorder, February. [74] Chen, C. C., and Raghavan, R. 1997. A Multiply-Fractured Horizontal Well in a Rectangular

Drainage Region. SPE J. 2 (December). 455-465. SPE 37072 PA. [75] Fekete reference material, 2015, Dual Porosity.

http://www.fekete.com/SAN/WebHelp/FeketeHarmony/Harmony_WebHelp/Content/HTML_Files/Reference_Material/General_Concepts/Dual_Porosity.htm, sited: May 1st 2015.

[76] Nobakht, M., Clarkson, C., and Kaviani, D. 2011. New Type Curves for Analyzing Horizontal Well with Multiple Fractures in Shale Gas Reservoirs. Paper CSUG/SPE 149397 presented at the Canadian Unconventional Resources Conference, Calgary, Alberta, Canada, 15-17 November.

[77] Rbeawi, S. A., and Tiab, D. 2013. Transient Pressure Analysis of Horizontal Wells in a Multi-Boundary System. American Journal of Engineering Research 2 (4): 44-66. DOI: 2320-0847.

[78] Ozkan, E., Brown, M., Raghavan, R., and Kazemi, H. 2009. Comparison of Fractured Horizontal –Well Performance in Conventional and Unconventional Reservoirs. Paper SPE 121290 presented at the 2009 SPE Western Regional Meeting, San Jose, California, USA, 24-26 March.

[79] Brown, M., Ozkan, E., Raghavan, R., and Kazemi, H. 2009. Practical Solutions for Pressure Transient Responses of Fractured Horizontal Wells in Unconventional Reservoirs. Paper SPE 125043 presented at the 2009 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, USA, 4-7 October.

[80] Stalgorova, E., and Mattar, L. 2012. Practical Analytical Model to Simulate Production of Horizontal Wells with Branch Fractures. Paper SPE 162515 presented at the SPE Canadian Unconventional Resources Conference, Calgary, Alberta, Canada, 30 October – 1 November.

67

Page 85: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[81] Daneshy, A. A. 2003. Off Balance Growth: A New Concept in Hydraulic Fracturing. Journal of Petroleum Technology 55(4): 78-85. SPE 80992-MS. http:dx.doi.org/10.2118/80992-MS.

[82] Arps, J. J. 1945. Analysis of Decline Curves. Trans. AIME, 160, 228. [83] Fetkovich, M. J. 1980. Decline Curve Analysis using Type Curves. JPT (June), 1065. [84] Blasingame, T. A., McGray, T. I., Lee, W. J. 1991. Decline Curve Analysis for Variable

Pressure Drop/Variable Flow rate Systems. Paper SPE 21513 presented at the SPE Gas Technology Symposium, 23-24 January.

[85] Wattenbarger, R. A., El-Banbi, A. H., Villegas, M. E., and Maggard, J. B. 1998. Production Analysis of Linear Flow into Fractured Tight Gas Wells. Paper SPE 39931 presented at the 1998 Rocky Mountain Regional/Low Permeability Reservoirs Symposium and Exhibition, Denver, USA, 5-8 April.

[86] Agarwal, R. G., Gardner, D. C., Kleinsteiber, S. W. and Fussell, D. D. 1998. Analyzing Well Production Data Using Combined Type Curve and Decline Curve Concepts. Paper SPE 57916 presented at the 1998 SPE Annual Technical Conference and Exhibition, New Orleans, 27-30 September.

68

Page 86: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

CHAPTER 3: ESTIMATION OF FRACTURE

CHARACTERISTIC WITHIN STIMULATED ROCK

VOLUME USING FINITE ELEMENT AND SEMI-

ANALYTICAL APPROACHES

Summary Hydraulic fracturing has been proven to effectively increase the drainage area and the permeability

of unconventional oil and gas reservoirs by creating a fracture network or stimulated rock volume

(SRV) within the reservoir rock. The dimensions of the SRV and its permeability are the key

parameters that enhance the unconventional reservoirs’ performance after the hydraulic fracture

operation. Simulation of the SRV to obtain its dimensions and permeability is required to

determine the optimum hydraulic fracture treatment parameters and production. In this study, finite

element analysis is used to determine the SRV characteristics based on field data from a hydraulic

fracturing job in a horizontal well penetrating the Glauconitic formation in the Hoadley Field,

Alberta, Canada. The dimensions of the SRV are calibrated from microseismic data. The fracture

propagation pressure of the finite element model is matched to the field value by altering the

permeability of the SRV. The matched model is used to obtain the in-situ stress changes and the

pressure drop within the SRV. The SRV permeability and the pressure drop are used to calculate

the aperture, the number and the spacing of the fractures within the SRV using a semi-analytical

approach. The final outputs can be used to optimize the future hydraulic fracture design at the

Hoadley field or at other fields that have similar geomechanical properties. It could also be used

to provide estimates of changes in the in-situ stresses around a stimulated horizontal wellbore.

3.1 Introduction

With declining conventional fossil fuel production, there has been greater production from

unconventional resources such as tight gas reservoirs. With immense petroleum volume in place

and long-term production potential, tight gas has become a crucial component of the fossil fuel

energy future. These unconventional reservoirs require stimulation technologies, such as hydraulic

69

Page 87: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

fracturing, to be commercially productive. Multistage hydraulically fractured horizontal wells are

becoming the standard method to produce unconventional gas reservoirs. Multi-fractured

horizontal well completion in a naturally fractured reservoir results in a fracture network or

stimulated rock volume (SRV) with large drainage area and high permeability created by the

interaction of the hydraulic fractures and the natural fractures. The dimensions and the

permeability of the SRV are the main factors that control the performance of the recovery process

from unconventional reservoirs. Despite its importance, the interaction between hydraulic

fractures and natural fractures have not been fully understood due to difficulties and expensive

cost of interpreting core (laboratory) and microseismic (field) data.

3.1.1 Objective of Study

Previous studies have shown that SRV permeability is affected by important parameters

including hydraulic fracture injection operating conditions, in-situ stresses, stimulated rock

volume dimensions and natural fracture characteristics.

In this study, a three-dimensional (3D) finite element analysis (FEA) is used to model the fluid

injection and pressure into the SRV in a tight gas formation. The SRV is assumed to be isotropic

linear elastic with an enhanced permeability. The simulated effective permeability from the FEA

is used as an input into a semi-analytical approach to determine the fracture network characteristics

such as fracture aperture and spacing. The semi-analytical approach makes uses of equivalent flow

characteristics and the mass conservation principle.

The objectives of this study are: (i) determination of SRV permeability during hydraulic

fracturing using 3D FEA and (ii) development of a semi-analytical approach to evaluate the

fracture characteristics within the SRV.

70

Page 88: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.2 Literature Review

3.2.1 Prediction of Stimulated Rock Volume Permeability

Oda (1986) predicted the effective permeability tensor of a naturally fractured reservoir using

the geometry of the fracture network [1]. Oda (1986) used cubic law to model the reservoir as a

cubic block of a rock containing the natural fractures that were arranged into smaller cubes [1].

Oda (1986) found that the effective permeability could be determined using the total number of

cubes and fracture intersections, the fracture intersection length and aperture, the fluid viscosity,

and the distance between the two adjacent cubes [1].

Rahman et al. (2002) proposed the SRV permeability as a function of the stimulation pressure,

the in-situ stresses, and the fracture density [2]. A model that considered both the fracture

propagation and the shear slippage of natural fractures was developed [2]. Rahman et al.’s (2002)

study results showed that the average permeability of the SRV increased sharply with a increase

stimulation pressure beyond a threshold value [2]. The threshold pressure was a function of the in-

situ stresses and the natural fractures properties in the reservoir [2]. Rahman et al. (2002) also

found that the permeability enhancement was nearly linearly dependent on the natural fracture

density [2].

Ge and Ghassemi (2011) developed a procedure to determine the SRV permeability by

matching the calculated SRV dimensions with the volume from microseismic monitoring [3]. Ge

and Ghassemi (2011) proposed a relationship between the net fracture pressure with the SRV

permeability and the SRV dimensions. A permeability was initially guessed, and then for the

selected net fracture pressure, the pore pressure and in-situ stress distributions were found [3]. This

pore pressure distribution was used with a Mohr-Coulomb failure envelope to determine the

amount of the shear failure in the formation around the hydraulically induced fracture [3]. From

the failure envelope, the extent of the zone of shear failure around the hydraulic fracture (thus

defining the SRV) was constructed [3]. The trial and error procedure were repeated to predict the

SRV permeability until the SRV dimensions matched the microseismic SRV [3].

Bahrami et al. (2012) proposed a semi-analytical equation to model the fracture permeability

as a function of the well test permeability, the fracture aperture, and the fracture spacing [4]. A

71

Page 89: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

fitted correlation was derived based on the sensitivity analysis of the well test permeability with

the fracture permeability, aperture, spacing and compressibility [4].

Johri and Zoback (2013) conducted a study that showed the fracture permeability was

enhanced during the hydraulic fracturing caused by the slip of natural fractures [5].

Nassir (2013) developed a coupled reservoir-geomechanics model to determine the SRV

permeability [6]. The maximum permeability enhancement was observed to result from high

injection rates [6].

3.2.2 Hydraulic Fracture - Natural Fracture Interaction

Interactions between hydraulic fractures and natural fractures have been studied both

experimentally and numerically since the 1980s. Hydraulic fracturing experiments were conducted

by Blanton (1982) using a pre-fractured material in the laboratory under tri-axial loads [7]. His

experiments results revealed that the hydraulic fractures preferred to cross the pre-existing fracture

only under high differential stresses and high angles of approach [7].

Warpinski and Teufel (1987) showed that the interaction of hydraulic fractures and natural

fractures was affected by the in-situ stress contrast, the natural fracture spacing and permeability,

and the hydraulic fractures treatment pressure [8].

Shimizu et al. (2014) investigated the influence of natural fracture permeability and approach

angle on the hydraulic fracturing simulation [9]. It was found that a high approach angle and a low

natural fracture permeability caused the hydraulic fracture to ignore the existence of the natural

fractures [9]. And the hydraulic fracture propagated straight to the direction of the maximum in-

situ stress [9].

Pirayehgar and Dusseault (2014) used the Universal Distinct Element Code (UDEC) to analyze

the fluid injection through the fractures in an impermeable reservoir [10]. Pirayehgar and

Dusseault’s (2014) results showed branching was occurring when the stress ratio was small [10].

Branching was a condition in which no explicitly favored path existed for fractures to propagate

[10]. Their simulation showed that branching occurred at a short distance from the injection point

and the branching was usually suppressed under a high in-situ stress ratio [10].

72

Page 90: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.2.3 Estimation of Pressure Drop and Fracture Aperture

Barenblatt et al. (1960) modeled the pressure drop in the naturally fractured reservoir by using

a high permeability fracture and a low permeability porous rock [11]. From their study, a

relationship was found between liquid pressure in the matrix and the natural fractures [11].

Another study was performed by Kim and Schecter (2009), who used a discrete fracture

network model, image logs and core analysis to estimate fracture aperture [12]. Their model results

showed that the fracture aperture could be estimated by incorporating outcrop maps, image logs,

computed tomography imaging, and core fracture data. Yu and Aguilera (2012) developed an

analytical model to solve a 3D pressure diffusion equation to predict the pressure drop within the

SRV [13]. They determined the hydraulic diffusivity coefficient to calibrate the model and then

the pressure drop within the SRV was predicted [13].

Izadi and Elsworth (2014) performed a study combining the cubic law and the Darcy’s law to

determine the pressure drop within the SRV [14]. This combined relationship was a function of

the initial fracture aperture, permeability, and spacing [14].

Estimation of fracture network characteristics such as the fracture aperture and spacing have

not been modeled adequately in recent studies.

3.2.4 Rock Geomechanical Properties Effect on Stimulated Rock Volume

Valko and Economides (2001) stated that the key geomechanical properties used in the

analytical models (KGD and PKN) for the hydraulic fracture are Young's modulus and Poisson’s

ratio [15].

Bratton (2011) showed that in-situ stress anisotropy dictated the complexity of the fracture

network where smaller in-situ stress anisotropy would not produce a pronounced directional

preference of the hydraulic fracture network [16].

Nassir (2013) showed that the cohesion was an important parameter in determining the Mohr-

Coulomb shear failure envelope; where the effective stresses were proportional to the cohesion of

the rock in the Mohr-Coulomb shear failure criterion [6]:

73

Page 91: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

𝜎𝜎1′ = 2𝐶𝐶 𝐶𝐶𝑜𝑜𝐶𝐶𝜙𝜙1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙

+ 𝜎𝜎3′1+𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙

(1)

where 𝜎𝜎1′ is the maximum effective stress, 𝜎𝜎3′ is the minimum effective stress, C is the cohesion

and 𝑡𝑡 is the friction angle.

Fang et al. (2015) found a relationship between the important parameters such as the bulk

modulus, the fracture aperture, spacing, and permeability [17].

3.3 Hoadley Field Properties

The part of the Hoadley Field in Rimbey, Alberta, Canada focused on in this study has a target

formation of tight gas hosted in a Glauconitic Formation interval located at 1892 m TVD with a

thickness of about 43 m. The stress regime in this field is a strike slip fault regime where the

maximum in-situ stress is the maximum horizontal stress with direction of 48o NE, the intermediate

in-situ stress is the vertical stress, and the minimum in-situ stress is the minimum horizontal stress.

The overlying formations are the Medicine River coal formation with a thickness of about 5 m and

the Mannville sandstone formation with a thickness of about 80 m. The underlying formation is

the Ostracod limestone formation with a thickness of about 87 m. At the initial in-situ condition,

the Glauconitic Formation average permeability is 0.07 mD (it was determined from build-up test

in the study well and nearby well core test [18]). Natural fractures were observed from the

microseismic data and inferred from the Mohr-Coulomb failure envelope with a predicted inclined

angle of 30o respect to the maximum horizontal stress. The initial formation properties are listed

in Table 3.1. The dynamic modulus properties are calculated from logs. It is assumed that the logs

represent intact rock nearby wellbore with the dynamic and the static modulus properties are

assumed to be equal. And they are expected to be higher compared to static modulus as proposed

by Jizba et al. (1990) [19]. The results of the estimated SRV dimensions from the microseismic

monitoring are listed in Table 3.2.

74

Page 92: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 3.1: Initial formation properties.

Initial formation properties Values

Pore pressure gradient (kPa/m) 4.86

Pore pressure at injection depth (MPa) 9.19

Total minimum horizontal stress gradient (kPa/m) 11.66

Total minimum horizontal stress at injection depth (MPa) 22.06

Total maximum horizontal stress gradient (kPa/m) 25.79

Total maximum horizontal stress at injection depth (MPa) 48.78

Total vertical stress gradient (kPa/m) 24.09

Total vertical stress at injection depth (MPa) 45.57

Initial permeability (mD) 0.07

Dynamic Poisson ratio of Mannville Formation 0.24

Dynamic Poisson ratio of Medicine River Coal Formation 0.28

Dynamic Poisson ratio of Glauconitic Formation 0.23

Dynamic Poisson ratio of Ostracod Formation 0.20

Dynamic Young's modulus of Mannville Formation (GPa) 35.38

Dynamic Young's modulus of Medicine River Coal Formation (GPa) 5.48

Dynamic Young's modulus of Glauconitic Formation (GPa) 45.04

Dynamic Young's modulus of Ostracod Formation (GPa) 45.03

Table 3.2: SRV dimensions.

SRV Dimension Values

SRV length (m) 174

SRV width (m) 60

SRV height (m) 60

90% SRV length (m) 157

90% SRV width (m) 54

90% SRV height (m) 54

75

Page 93: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.4 Finite Element Analysis Model

3.4.1 Constitutive Model for Tight Sand in Finite Element Analysis

Tight sandstone is dominated by clean sandstone deposited in high-energy depositional settings

whose intergranular pores have been largely occluded by authigenic cements (mainly quartz and

calcite). Post-depositional diagenetic events reduce the effective porosity and permeability of the

rock. Masters (1979) defined low permeability gas saturated Cretaceous sandstone reservoirs of

western Alberta characteristics as: 1. low porosity (value in range 7-15%), 2. low permeability

(value in range 0.1-1 mD), and 3. moderate water saturation (value in range 34-45%) [20].

This research treats the tight sandstone as a linear elastic material and porous medium. The

constitutive equation for the material is given by:

𝜎𝜎 = 𝐷𝐷𝑒𝑒𝑒𝑒𝜀𝜀𝑒𝑒𝑒𝑒 (2)

where σ is the total stress, Del is the fourth order elasticity tensor, and εel is the total elastic strain.

The simplest form of the linear elasticity is the isotropic case, where the stress-strain relationship

is given by (Abaqus 2013) [21]:

⎩⎪⎨

⎪⎧𝜀𝜀11𝜀𝜀22𝜀𝜀33𝛾𝛾12𝛾𝛾13𝛾𝛾23⎭

⎪⎬

⎪⎫

=

⎣⎢⎢⎢⎢⎢⎢⎢⎢⎡1𝜋𝜋

−𝑣𝑣𝜋𝜋

−𝑣𝑣𝜋𝜋

0 0 0−𝑣𝑣𝜋𝜋

1𝜋𝜋

−𝑣𝑣𝜋𝜋

0 0 0

−𝑣𝑣𝜋𝜋000

−𝑣𝑣𝜋𝜋000

1𝜋𝜋

0 0 0

0 1𝐺𝐺 0 0

0 0 1𝐺𝐺

0

0 0 0 1𝐺𝐺 ⎦⎥⎥⎥⎥⎥⎥⎥⎥⎤

⎩⎪⎨

⎪⎧𝜎𝜎11𝜎𝜎22𝜎𝜎33𝜎𝜎12𝜎𝜎13𝜎𝜎23⎭

⎪⎬

⎪⎫

(3)

where E is Young’s modulus, ν is Poisson’s ratio, G is shear modulus, ε is normal strain, and γ is

shear strain. The porous medium model used here considers the reservoir rock filled with a single

fluid phase (linear linked gel fracturing fluid of WF130 with viscosity of 30 cP and specific gravity

of 1.028). It assumes that the constitutive response of the porous medium follows the linear

elasticity for the liquid and the solid together with a constitutive theory for the solid skeleton as

76

Page 94: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

explained by (Nuth and Laloui 2008) [22]. Temperature effects are considered as small and thus

the system is considered as isothermal. Biot’s coefficient is assumed to be unity.

3.4.2 Pore Fluid Flow in Finite Element Analysis

For fluid flow in a porous medium, the governing equation is the Forchheimer’s modification

of the Darcy’s law that relates the fluid flow velocity to the pressure as follows [21]:

𝑐𝑐𝑡𝑡𝑣𝑣𝑤𝑤 + 𝛽𝛽𝑐𝑐𝑡𝑡(𝑣𝑣𝑤𝑤)2 = −𝑘𝑘� 1𝑔𝑔𝜌𝜌𝑤𝑤

�𝜕𝜕𝑢𝑢𝑤𝑤𝜕𝜕𝜕𝜕

− 𝑔𝑔𝜌𝜌𝑤𝑤𝜕𝜕𝑧𝑧𝜕𝜕𝜕𝜕� (4)

𝛽𝛽 = 2.33 1010

𝑘𝑘1.201 (5)

𝐾𝐾� = 𝜐𝜐𝑔𝑔𝑘𝑘� (6)

where sn is the fluid volume fraction, β is the Forchheimer’s coefficient, vw is the fluid velocity,

uw is the wetting liquid pore pressure, x is the position, ρw is the fluid density, g is the gravitational

acceleration, 𝑘𝑘� is the hydraulic conductivity in m/s (permeability used in the FEA), 𝐾𝐾� is the

permeability in m2, and 𝜐𝜐 is the wetting liquid kinematic viscosity.

3.4.3 Finite Element Analysis Model

The commercial finite element software package, Abaqus [9], is used to model the interaction

of fluid and solid during the fracture fluid injection. The code uses the Petrov-Galerkin finite

element method to solve the governing equations for both solid (stress equilibrium) and fluid

phases (continuity). The liquid is injected into the domain through a boundary condition at the

location of the well injection port.

Within the porous medium, the flow of liquid obeys the Forchheimer's modification of the

Darcy’s law. For the fluid flow in the porous medium, a linear tetrahedral finite element (equal

order for velocity and pressure) is used [9]. For the solid mechanics (displacement and pore

pressure), a C3D4P four-node linear tetrahedral finite element is used [9]. The governing equations

are integrated through time using backward Euler time stepping. At each time step, the set of

coupled non-linear equilibrium and continuity equations are solved using Newton's method.

77

Page 95: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.4.4 Model Geometry

The model, displayed in Figure 3.1, consists of a four-layer block computational domain with

the following overall dimensions: length in x-axis 240 m and width in y-axis 240 m. The total

block thickness is 215 m which includes the lower Ostracod formation with a thickness of 87 m,

the target sandstone formation (Glauconitic) interval with a thickness of 43 m, the thin layer of

coal (Medicine River) with a thickness of 5 m, and the top sandstone formation (Mannville

formation) with a thickness of 80 m. The bottom surface of the domain is located at a depth of

2,000 m. At a depth of 1,892 m, the hydraulic injection port is modeled as an open flow boundary

condition (the wellbore is not modeled). From the microseismic observations, the SRV half-length

in x-axis direction is 87 m, the SRV half-width in y-axis direction is 30 m, and the SRV total height

in z-axis direction is 60 m [23]. The model dimensions were chosen to be three times of the SRV

length to be representative of the stress field. The mesh sizes were chosen to be bigger size for

non-SRV (16 m) and smaller size for SRV (5 m). The mesh was generated by assigning elements

numbers for each edge.

Fig. 3.1: Finite element model mesh for hydraulic fracturing (SRV) simulation.

78

Page 96: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The model is partitioned into seven intervals: the Glauconitic-non SRV, the Glauconitic-SRV,

the Medicine River-non SRV, the Medicine River-SRV, the Mannville-non SRV, the Mannville-

SRV and the Ostracod-non SRV. The intervals are derived from the examination of the

microseismic SRV dimensions. The model assumes that the SRV is a relatively high permeability

volume during the fluid injection. Each layer of the model has its own density, Young's modulus,

Poisson's ratio, permeability, void ratio and specific weight of the host fluid. The Young’s modulus

and Poisson’s ratio were derived from logs.

Because of the inherent symmetry of the geometrical model about the x=0 and y=0 planes,

only one quarter of the geometrical model is meshed and analyzed. The meshed model is

discretized into 9,676 tetrahedral finite elements and the meshed model consists of two distinct

sections: (i) the bulk section representing the four formations (with mesh size equal to 16 m in

length, height and width) and (ii) the finer section covering the expected SRV including the

hydraulic fracture injection port (with mesh size 5 m in length, height and width).

3.4.5 Initial and Boundary Conditions

To specify the initial condition (in equilibrium state), the parameters that need to be defined

are: (a) the effective stresses with the maximum in-situ stress being the maximum horizontal stress

which is parallel to the direction of hydraulic fracture, (b) the pore pressure, and (c) the void ratio.

The initial properties of the formation are listed in Table 3.1.

To achieve the initial static equilibrium, the loads on the domain are: (a) the gravity load is

applied on the whole model with applied acceleration 9.8 m/s2 in the negative z-direction and (b)

the overburden stress applied to the top surface is equal to 43 MPa (assumed uniform across the

top of the model domain).

In the second step, a single hydraulic fracture stage is simulated. The hydraulic fracture

operation is divided into three steps with total duration of 37.5 minutes (2,250 s). The durations of

the three steps are 1, 10, and 2,239 s. The first two steps are divided into 10 equal sub-steps and

the last step has time steps equal to 5 s. The reason for choosing smaller time increments in the

early stages is to capture the consolidation that occurs following the load application. In the second

step there is an additional load applied which is the injection velocity based on 5 m3/min injection

79

Page 97: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

flow rate. There is also a pore pressure boundary condition on the four outer surfaces that is equal

to the initial pore pressure.

In total, an average of 90 minutes of (wall-clock) time is required to simulate the model on a

4-core, 2.2 GHz machine with 16 GB of memory.

3.5 Results and Discussion

3.5.1 Determination of Effective Permeability

To determine the effective permeability of the SRV, a search was conducted by varying the

values of the permeability until the average simulated fracture propagation pressure matched the

field data.

The value of the pressure to be matched, from the field average fracture propagation pressure

data, is equal to 27.62 MPa as shown in Figure 3.2. This analysis does not model the deformation

and the fracture breakdown pressure. Therefore the model is only matching the average fracture

propagation pressure at the final injection time. From the search, the effective permeability that

yields the best match to the field data is found to be equal to 23.4 D (permeability is assumed to

be isotropic) for 100% SRV dimensions (Case 1).

80

Page 98: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 3.2: Bottom-hole injection pressure (field) data and predicted results from FEA.

To find the effect of the SRV dimensions change on the effective permeability and to deal with

the microseismic uncertainties, simulations are done for Case 2 (90% of base SRV dimensions).

For Case 2, each SRV dimension is reduced by 10% resulting in a reduction of the SRV volume

by 33%. The matched effective permeability for Case 2 is equal to 45.8 D.

Figure 3.2 also shows that Case 2 has a higher bottom-hole injection pressure gradient (6.6

kPa/s) between early and late time (from 0 second to 2,002 s) compared to Case 1 (4.9 kPa/s).

However, Case 1 and Case 2 have similar bottom-hole injection pressure gradients at late time

(from 2,002 s until the final injection time). Case 3 is simulated by reducing the grid dimensions

to 63% of Case 1. It yields 0.17 % change in the results of the final pore pressure as shown in

Figure 3.2 (0.7% change in the deformation and 0.1% change in the stresses, they are not shown

in Figure 3.2). Figure 3.3 shows the pore pressure distribution within and around the SRV versus

time for Case 1. Figure 3.3a is the pre-calculation pore pressure distribution (initial condition) and

Figure 3.3b is the initial pore pressure distribution after the in-situ stresses are applied on the

domain. Figures 3.3c, 3.3d, and 3.3e display the pore pressure distributions after 1, 551, and 1,101

s of the fluid injection. After 1,101 s of fluid injection, the pore pressure has almost reached the

target pressure. Finally, Figure 3.3f shows the maximum pore pressure after 2,250 s at the injection

7

14

21

28

35

42

0 500 1000 1500 2000 2500

BHP

(MPa

)

Time (s)

Bottom Hole Pressure

BHP Field Data BHP FEA for Case 1BHP FEA for Case 2 BHP FEA for Case 3

81

Page 99: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

port and nearby the injection port is 27.62 MPa and the pore pressure on the SRV boundary is

equal to about 20.1 MPa (the pressure drop along the SRV is 7.52 MPa).

Fig. 3.3: Pore pressure (Pa) in different steps: (a) initial condition, (b) after in-situ stresses and boundary conditions are loaded on the domain step, (c) pump step-1 second, (d) pump step-551 s, (e) pump step-1,101 s, and (f) end of injection-2,250 s for Case 1 with k=23.4 D (deformation scale factor of 3,505.7 with final displacement of 6.853e-3 m).

Figure 3.4 shows that a uniform target bottom-hole pressure distribution within the SRV is

achieved faster in Case 2 compared to Case 1 as shown in the Figure 3.3f and Figure 3.4f.

82

Page 100: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 3.4: Pore pressure (Pa) in different steps: (a) initial condition, (b) after in-situ stresses and boundary conditions are loaded on the domain step, (c) pump step-1 second, (d) pump step-551 s, (e) pump step-1,101 s, and (f) end of injection-2,250 s for Case 2 with k=45.85 D (deformation scale factor of 3,016.07 with final displacement of 7,957e-3 m).

3.5.2 Parametric Studies of Young’s Modulus

Parametric studies were conducted with the reduction of the base Young’s modulus of the

Glauconitic, the Medicine River and the Mannville formations that form the SRV in the finite

element model and are in contact with the hydraulic fracture fluid. The Young's modulus of the

SRV in the finite element model was reduced from the base values to 90, 80, and 70% of the base

values. The results of the sensitivity runs for Case 1 show that the final injection pressures are

27.01, 26.39, and 25.76 MPa when the Young's modulus of the SRV are reduced to 90, 80, and

70%, respectively (Figure 3.5a). For Case 2 (Figure 3.5b), show that the final injection pressures

are 26.78, 25.97, and 25.14 MPa when the Young’s modulus of the SRV are reduced to 90, 80,

and 70%, respectively.

83

Page 101: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.5: Effect of Young’s modulus on bottom-hole pressure for (a) Case 1 and (b) Case 2.

Additional parametric study was also done to find the matched effective permeability with the

reduced values of the Young's modulus that led to an injection pressure of 27.62 MPa. The results

for Case 1 are as follows:

1. 70% of initial Young's modulus gives a matched effective permeability of 18.4 D,

2. 80% of initial Young's modulus gives a matched effective permeability of 20 D, and

3. 90% of initial Young's modulus gives a matched effective permeability of 21.5 D.

51015202530

0 500 1000 1500 2000 2500

Pore

Pre

ssur

e (M

Pa)

Time (s)

Young's modulus Effect on Pore Pressure - Case 1 (k=23.4 D)

Base E 90%E 80%E 70%E

51015202530

0 500 1000 1500 2000 2500

Pore

Pre

ssur

e (M

Pa)

Time (s)

Young's modulus Effect on Pore Pressure - Case 2 (k=45.85 D)

Base E 90%E 80%E 70%E

84

Page 102: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The parametric studies done on Case 2 yield the following results:

4. 70% of initial Young's modulus gives a matched effective permeability of 27 D,

5. 80% of initial Young's modulus gives a matched effective permeability of 35.8 D, and

6. 90% of initial Young's modulus gives a matched effective permeability of 39.78 D.

From the parametric studies, the results show that the reduction of the Young's modulus lowers

the effective permeability of the SRV. Also, reduction in the Young’s modulus is not linearly

related with the reduction of the effective permeability. This permeability dependence with

Young’s modulus occurs due to the inverse relationship between Young’s modulus and fracture

aperture. Reduction in the SRV dimensions leads to a higher increase of the effective permeability

in a lower Young’s modulus.

3.5.3 Pore Pressure and In-Situ Stresses from Finite Element Analysis

The pore pressure and the total maximum horizontal stresses are plotted versus distance (in the

SRV length direction) for Case 1 at the base Young's modulus in Figure 3.6. The results reveal

that the maximum pressure decrease happens at a distance of about 10 m from the injection port

and the pressure drop from the 10 m distance to the SRV boundary is small.

85

Page 103: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.6: (a) Pore pressure and (b) total maximum horizontal stress as a function of distance along SRV length at various injection times for Case 1 with k=23.4D (assuming Biot’s constant=1).

The maximum horizontal stress increase is within a distance of 10 m from the injection port.

This stress increases due to the poro-elastic effect and the fluid injection. The causes of increase

in the total in-situ stresses are explained by Vermylen (2011) [24]. Vermylen (2011) stated that

there were three general causes of the stress changes in the reservoir due to the hydraulic fracture:

(a) creation of tensile fractures of induced hydraulic fracture, (b) poroelastic effects when the fluid

0

10

20

30

0 10 20 30 40 50 60 70 80 90Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Length for Case 1 (k=23.4D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

-65

-60

-55

-50

-450 10 20 30 40 50 60 70 80 90

Tota

l SH

(MPa

)

Distance from injection (m)

Total SH along SRV Length for Case 1 (k=23.4D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

86

Page 104: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

leaked off from the hydraulic fracture and increased the pore pressure in the reservoir, and (c)

propagation of hydraulic fractures [24]. Afterward the stress would relieve and cause stress

relaxation [24]. This combination of actions would change the in-situ stresses [24]. The total

minimum horizontal stress and the total vertical stress also produce responses similar to the

maximum horizontal stress as shown in Figure 3.7 and Figure 3.8, respectively.

The total maximum stresses drop is small from a distance of about 10 m to the SRV boundary

as shown in Figure 3.6. The total maximum horizontal stress after the hydraulic fracture increases

to between 49 MPa and 61 MPa (in-situ total maximum horizontal stress is 49 MPa) near the

injection port. Then, there are no changes between the distances of 10 m from the injection port to

the SRV boundary over the time period of 1 second. However, at late injection (551, 1,101 and

2,250 s), there are smaller total maximum horizontal stress increases to between 50 MPa and 56.5

MPa between distance of 10 m to the SRV boundary. The changes in the total in-situ stresses

calculated by Abaqus might be overestimated because in this study, the Biot’s constant is assumed

to be unity. This assumption yields that a change in the pore pressure generates an equal and

opposite change in the total in-situ stresses, according to the principle of effective stress. For Biot’s

constant of less than unity which is valid for the Glauconitic formation, the changes in the total in-

situ stresses induced by the pore pressure would be reduced.

87

Page 105: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.7: (a) Pore pressure and (b) total minimum horizontal stress as a function of distance along SRV width at various injection times for Case 1 with k=23.4D (assuming Biot’s constant=1).

05

1015202530

0 5 10 15 20 25 30 35Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Width for Case 1 (k=23.4D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

-35

-30

-25

-200 5 10 15 20 25 30 35

Tota

l Sh

(MPa

)

Distance from injection (m)

Total Sh along SRV Width for Case 1 (k=23.4D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

88

Page 106: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.8: (a) Pore pressure and (b) total vertical stress as a function of distance along SRV height above injection ports at various injection times for Case 1 with k=23.4D (assuming Biot’s constant=1).

The pore pressure and the total in-situ stresses for Case 2 as shown in Figure 3.9, Figure 3.10

and Figure 3.11) produce responses similar to those in Case 1. The only differences between Case

1 and Case 2 results are Case 1 produces the pore pressure of 20.1 MPa at the SRV boundary with

the pressure drop along the SRV length of 7.52 MPa where Case 2 produces higher pore pressure

on the SRV boundary of 23.84 MPa and lower pressure drop along the SRV length of 3.84 MPa.

05

1015202530

0 5 10 15 20 25 30 35 40Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Top Height for Case 1 (k=23.4D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

-60

-55

-50

-45

-400 5 10 15 20 25 30 35 40

Tota

l SV

(MPa

)

Distance from injection (m)

Total SV along SRV Top Height for Case 1 (k=23.4D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

89

Page 107: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The other differences are Case 2 results at the final injection time (2,250 s) showing similar stresses

gradient with Case 1 at shorter distance and higher gradient at longer distance from the injection

port. The final horizontal stresses at the SRV boundaries during final injection time for Case 2 are

higher compared to Case 1.

(a)

(b)

Fig. 3.9: (a) Pore pressure and (b) total maximum horizontal stress as a function of distance along SRV length at various injection times for Case 2 with k=45.85D (assuming Biot’s constant=1).

05

1015202530

0 10 20 30 40 50 60 70 80 90Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Length for Case 2 (k=45.85D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

-65

-60

-55

-50

-450 10 20 30 40 50 60 70 80 90

Tota

l SH

(MPa

)

Distance from injection (m)

Total SH along SRV Length for Case 2 (k=45.85D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

90

Page 108: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.10: (a) Pore pressure and (b) total minimum horizontal stress as a function of distance along SRV width at various injection times for Case 2 with k=45.85D (assuming Biot’s constant=1).

05

1015202530

0 5 10 15 20 25 30Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Width for Case 2 (k=45.85D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

-30

-25

-200 5 10 15 20 25 30

Tota

l Sh

(MPa

)

Distance from injection (m)

Total Sh along SRV Width for Case 2 (k=45.85D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

91

Page 109: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.11: (a) Pore pressure and (b) total minimum horizontal stress as a function of distance along SRV height above injection ports at various injection times for Case 2 with k=45.85D (assuming Biot’s constant=1).

05

1015202530

0 5 10 15 20 25 30 35Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Height for Case 2 (k=45.85D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

-60-55-50-45-40

0 5 10 15 20 25 30 35

Tota

l SV

(MPa

)

Distance from injection (m)

Total SV along SRV Height for Case 2 (k=45.85D)

(t=1s) (t=551s) (t=1101s) (t=2250s)

92

Page 110: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.5.4 Determination of Fracture Aperture Using Cubic Law

The pore pressure gradient along a fracture is calculated analytically using the cubic equation

as a function of fracture aperture and half-length [25].

𝑄𝑄 = −𝐿𝐿 𝑑𝑑3

12𝜇𝜇𝑑𝑑𝑃𝑃𝑑𝑑𝜕𝜕

(7)

where Q is flow rate, L is fracture width in direction normal to fluid flow, d is fracture aperture, µ

is fluid viscosity, P is fluid pressure, and x is distance.

The SRV half-length and the fracture aperture are varied as shown in Figure 3.12a for Case 1

and Figure 3.12b for Case 2. For Case 1 (Figure 3.12a), the fracture aperture is varied from 1.43

mm to 7 mm with the maximum SRV half-length of 87 m. The fracture aperture of 1.43 mm

matches the finite element pore pressure of 20.1 MPa at the SRV boundary. Figure 3.12b also

shows that fracture aperture is varied from 1.78 mm to 7 mm with the maximum SRV half-length

of 78 m for Case 2. The fracture aperture of 1.78 mm matches the finite element pore pressure of

23.84 MPa at the SRV boundary.

93

Page 111: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.12: Pore pressure as a function of distance using the cubic law equation for (a) Case 1 and (b) Case 2.

The differences between the pore pressure modeling of cubic law model and finite element are

that the cubic law model assumes a steady state flow whereas the finite element modeling assumes

transient flow and the cubic equation does not take geomechanics effect into account.

2022242628

0 10 20 30 40 50 60 70 80 90Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Length -Cubic Law for Case 1

(d=1.43mm) (d=2mm) (d=3mm)(d=4mm) (d=5mm) (d=6mm)(d=7mm)

232425262728

0 10 20 30 40 50 60 70 80 90Pore

Pre

ssur

e (M

Pa)

Distance from injection (m)

Pore Pressure along SRV Length -Cubic Law for Case 2

(d=1.78mm) (d=2mm) (d=3mm)(d=4mm) (d=5mm) (d=6mm)(d=7mm)

94

Page 112: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.5.5 Determination of Fracture Characteristics Using Semi-Analytical

Approach

In this section, a semi-analytical approach is developed to calculate the fracture characteristics.

The semi-analytical approach neglects leak-off. The fracture network within the SRV is assumed

to consist of the major and the minor fractures (the minor fractures are assumed to be the natural

fractures). The hydraulic fractures (major fractures) grow in the SRV length direction (x-axis) and

the natural fractures (minor fractures) are assumed to have an inclined angle of 30o respect to the

maximum horizontal stress (Figure 3.13). The major fractures contribute to the SRV enhanced

permeability (since major fractures connect SRV boundaries) and the injected volume where the

minor fractures are assumed only to contribute to the injected volume because the minor fractures

are assumed to be dead-end (do not connect SRV boundaries). The approach assumes that the

major fractures created by the hydraulic fractures intercept the natural fractures and they propagate

along the natural fractures and create another major fractures inside the SRV. Therefore the

hydraulic fractures and the natural fractures are assumed to be connected and contribute to injected

fracture fluid volume.

Fig. 3.13: Top view of the minor fractures (natural fractures) are assumed to be inclined 30o with the major fractures (hydraulic fractures) (not to scale).

95

Page 113: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The numbers of the major and minor fractures, aperture and spacing are calculated based on

two criteria (i) equivalent flow characteristics and (ii) mass conservation:

𝑘𝑘�𝑆𝑆𝑆𝑆𝑉𝑉𝐴𝐴𝑆𝑆𝑆𝑆𝑉𝑉 = ∑𝑘𝑘�𝑓𝑓𝑟𝑟𝑡𝑡𝐶𝐶𝑡𝑡𝑢𝑢𝑟𝑟𝑒𝑒𝐴𝐴𝑓𝑓𝑟𝑟𝑡𝑡𝐶𝐶𝑡𝑡𝑢𝑢𝑟𝑟𝑒𝑒 (7)

𝐼𝐼𝑡𝑡𝐼𝐼𝐼𝐼𝑐𝑐𝑡𝑡𝐼𝐼𝐼𝐼 𝑣𝑣𝑐𝑐𝑣𝑣𝑢𝑢𝑚𝑚𝐼𝐼 = 𝑓𝑓𝑟𝑟𝑡𝑡𝑐𝑐𝑡𝑡𝑢𝑢𝑟𝑟𝐼𝐼 𝑣𝑣𝑐𝑐𝑣𝑣𝑢𝑢𝑚𝑚𝐼𝐼 (8)

In Equation (7), 𝑘𝑘�𝑆𝑆𝑆𝑆𝑉𝑉 and ASRV are the effective permeability (from FEA) and flow area (SRV

width by SRV height) of the SRV, respectively. 𝑘𝑘�𝑓𝑓𝑟𝑟𝑡𝑡𝐶𝐶𝑡𝑡𝑢𝑢𝑟𝑟𝑒𝑒 and Afracture are the permeability (from

cubic law) and the flow area of each major fracture, respectively. Equation (8) assumes all the

injected volume is accumulated in the major and minor fractures as shown in Figure 3.13. The

procedures used to calculate the fracture aperture, numbers and spacing are shown in Figure 3.14.

Fig. 3.14: Procedures to calculate the fracture aperture, numbers and spacing.

2. Estimate 𝑘𝑘�𝑓𝑓𝑟𝑟𝑡𝑡𝐶𝐶𝑡𝑡𝑢𝑢𝑟𝑟𝑒𝑒𝐴𝐴𝑓𝑓𝑟𝑟𝑡𝑡𝐶𝐶𝑡𝑡𝑢𝑢𝑟𝑟𝑒𝑒 based on cubic law, Darcy’s law, and using major fracture aperture

3. Estimate major fracture number = 𝑘𝑘�𝑆𝑆𝑆𝑆𝑉𝑉𝐴𝐴𝑆𝑆𝑆𝑆𝑉𝑉 𝑘𝑘�𝑓𝑓𝑟𝑟𝑡𝑡𝐶𝐶𝑡𝑡𝑢𝑢𝑟𝑟𝑒𝑒𝐴𝐴𝑓𝑓𝑟𝑟𝑡𝑡𝐶𝐶𝑡𝑡𝑢𝑢𝑟𝑟𝑒𝑒�

4. Estimate total major fractures length = 𝑆𝑆𝑆𝑆𝑆𝑆 𝐿𝐿𝐼𝐼𝑡𝑡𝑔𝑔𝑡𝑡ℎ 𝑥𝑥 𝑡𝑡𝑢𝑢𝑚𝑚𝑛𝑛𝐼𝐼𝑟𝑟 𝑐𝑐𝑓𝑓 𝑚𝑚𝑡𝑡𝐼𝐼𝑐𝑐𝑟𝑟 𝑓𝑓𝑟𝑟𝑡𝑡𝑐𝑐𝑡𝑡𝑢𝑢𝑟𝑟𝐼𝐼

1. Calculate fracture aperture based on cubic law and using pressure drop and injection rate (from FEA)

7. Calculate minor fracture spacing= SRV lengthminor fracture number

, major fracture spacing = Hmajor fracture number

6. Calculate minor fracture number= 𝑇𝑇𝑐𝑐𝑡𝑡𝑡𝑡𝑣𝑣 𝑚𝑚𝑠𝑠𝑡𝑡𝑐𝑐𝑟𝑟 𝑓𝑓𝑟𝑟𝑡𝑡𝑐𝑐𝑡𝑡𝑢𝑢𝑟𝑟𝐼𝐼 𝑣𝑣𝐼𝐼𝑡𝑡𝑔𝑔𝑡𝑡ℎ � 𝑆𝑆𝑆𝑆𝑉𝑉 𝑤𝑤𝑖𝑖𝑑𝑑𝑡𝑡ℎ𝐶𝐶𝑜𝑜𝐶𝐶𝑖𝑖𝑛𝑛𝑒𝑒(90𝑜𝑜−𝛼𝛼)

�� , where α is angle

between major and minor fractures

5. Calculate total minor fracture length = 𝑇𝑇𝑐𝑐𝑡𝑡𝑡𝑡𝑣𝑣 𝑓𝑓𝑟𝑟𝑡𝑡𝑐𝑐𝑡𝑡𝑢𝑢𝑟𝑟𝐼𝐼 𝑣𝑣𝐼𝐼𝑡𝑡𝑔𝑔𝑡𝑡ℎ 𝑆𝑆𝑆𝑆𝑆𝑆 � 𝑉𝑉𝑖𝑖𝑤𝑤𝐻𝐻� −

𝑚𝑚𝑡𝑡𝐼𝐼𝑐𝑐𝑟𝑟 𝑓𝑓𝑟𝑟𝑡𝑡𝑐𝑐𝑡𝑡𝑢𝑢𝑟𝑟𝐼𝐼 𝑣𝑣𝐼𝐼𝑡𝑡𝑔𝑔𝑡𝑡ℎ, where Vi is injected volume, w is fracture aperture, H is fracture height

96

Page 114: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Case 1 produces a fracture aperture of 1.43 mm, a major fracture number of 5 with a spacing

of 13.1 m, and a minor fracture number of 12 with a spacing of 14.9 m. Case 2 produces a fracture

aperture of 1.78 mm, a major fracture number of 4 with a spacing of 13.1 m, and a minor fracture

number of 12 with a spacing of 12.9 m.

Figure 3.15a shows that for Case 1 with increasing a fracture aperture from 1 to 2.1 mm, the

number of major fractures and the pressure drop along the SRV length decrease where the number

of minor fracture increases until the fracture aperture is 1.43 mm and then the minor fracture

number decreases. This phenomenon is related to the major and minor fracture volumes because

the total injected volume is needed to match the total fracture volume from the major and the minor

fractures.

Figure 3.15b also shows that for Case 2 with increasing the fracture aperture from 1.1 to 2.25

mm, the number of major fractures and the pressure drop along the SRV length decrease where

the number of minor fractures increases until the fracture aperture is 2.1 mm and then the minor

fracture number decreases. The differences between Case 1 and Case 2 are for a fracture aperture

of 1.1 mm Case 1 yields 10 major fractures and Case 2 yields 18 major fractures.

97

Page 115: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 3.15: Relationship any number of fractures, fracture aperture and fracture pressure gradient for (a) Case 1 and (b) Case 2.

From Table 3.3, it can be seen that for constant fracture aperture and reducing the effective

permeability decreases the number of the major fractures and increases the number of minor

fractures. Case 2 produces responses similar to those in Case 1 with the constant fracture aperture.

050100150200250300

0

5

10

15

1 1.25 1.5 1.75 2 2.25

dP/d

x (k

Pa/m

)

Num

ber o

f Fra

ctur

es

Fracture Aperture (mm)

Number of Fractures for Case 1 (k=23.4D)

Number of Major Fracture Number of Minor Fracture

dP/dx

050100150200250

05

101520

1 1.25 1.5 1.75 2 2.25 2.5dP

/dx

(kPa

/m)

Num

ber o

f Fra

ctur

es

Fracture Aperture (mm)

Number of Fractures for Case 2 (k=45.85D)

Number of Major Fracture Number of Minor FracturedP/dx

98

Page 116: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 3.3: Parametric study results – effect of Young’s modulus and SRV on numbers of major and minor fractures.

Fracture aperture (mm)

Young’s modulus SRV

Permeability (D)

Number of Major Fracture

Number of Minor Fracture

1.43 100% Case 1 23.4 5 12

1.43 90% Case 1 21.5 4 12

1.43 80% Case 1 20 4 13

1.43 70% Case 1 18.4 4 13

Fracture width (mm)

Young's modulus SRV

Permeability (D)

Number of Major Fracture

Number of Minor Fracture

1.78 100% Case 2 45.85 4 12

1.78 90% Case 2 39.78 4 13

1.78 80% Case 2 35.8 3 13

1.78 70% Case 2 27 2 15

3.6 Conclusions

This study provides insights on fracture characteristics within a SRV created the hydraulic

fractures using finite element analysis and a semi analytical approach. The fracture characteristics

are the enhanced permeability, the pressure drop, and the in-situ stresses change created by the

hydraulic fractures within the SRV. The major and minor fracture aperture, numbers and spacing

are calculated using the semi-analytical approach with the inputs of the injected fracture fluid

volume and the simulated enhanced permeability. Case 1 produces the matched effective

permeability of 23.4 D, the fracture aperture of 1.43 mm, the major fracture number of 5 with the

spacing of 13.1 m, and the minor fracture number of 12 with the spacing of 14.9 m. Case 2 produces

the matched effective permeability of 45.89 D, the fracture aperture of 1.78 mm, the major fracture

number of 4 with the spacing of 13.1 m, and the minor fracture number of 12 with the spacing of

12.9 m. These parameters can be used to optimize the hydraulic fracture design (placement and

amount of injection) in the Glauconitic Formation at Hoadley field or at other fields that have

similar geomechanical properties. The results can be used to predict the reservoir production after

99

Page 117: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

the hydraulic fracture and provides how the in-situ stresses change around the stimulated

horizontal wellbore.

The developed workflow in this study using the finite element analysis and the semi-analytical

approach to characterize the fracture network within the SRV is novel. The conclusions of the

study are as follows:

1. Reduction of the Young’s modulus decreases the effective permeability.

2. Reduction of the SRV dimensions by 10% increases the effective permeability within the

SRV and decrease the pressure drop along the SRV length.

3. The hydraulic fracture induces an increase in the total in-situ stress values. The changes in

the total in-situ stresses calculated by Abaqus are likely overestimated.

4. Reduction of the SRV dimensions increases fractures aperture and number of major

fractures and decreases the number of minor fractures to produce the same pressure drop

along the SRV length.

5. Reduction of the Young’s modulus decreases the number of major fractures and increases

the number of minor fractures for constant fracture aperture.

100

Page 118: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

3.7 References

[1] Oda, M. 1986. An Equivalent Continuum Model for Coupled Stress and Fluid Flow Analysis in Jointed Rock Masses. Water Resource Research 22 (13): 1845-1856. http://dx.doi.org/10.1029/WR022i013p01845

[2] Rahman, M. K., Hossain, M. M., and Rahman, S. S. 2002. A Shear-Dilation Based Model for Evaluation of Hydraulically Stimulated Naturally Fractured Reservoirs. Int J Numer and Anal Meth Geomech 26 (5):469-497. http://dx.doi.org/10.1002/nag.208

[3] Ge, J. and Ghassemi, A. 2011. Permeability Enhancement in Shale Gas Reservoirs after Stimulation by Hydraulic Fracturing. Presented at the 45th US Rock Mechanics/Geomechanics Symposium, San Francisco, CA, 26-29 June. ARMA 11-514.

[4] Bahrami, H., Rezae, R., and Hossain, M. 2012, Characterizing Natural Fractures Productivity in Tight Gas. J Pet Explor Prod Technol 2 (26):107-115.

http://dx.doi.org/10.1007/s13202-012-0026-x [5] Johri, M., and Zoback, M. D. 2013. The Evolution of Stimulated rock volume during

Hydraulic Stimulation of Shale Gas Formations. Presented at the Unconventional Resources Technology Conference, Denver, Colorado, 12-14 August. URTeC 1575434.

[6] Nassir, M. 2013. Geomechanical Coupled Modeling of Shear Fracturing in Non-Conventional Reservoirs. Ph. D. Thesis, University of Calgary, Calgary, Alberta (January 2013).

[7] Blanton, T. L. 1982. An Experimental Study of Interaction between Hydraulically Induced and Pre-existing Fractures. Presented at the SPE/DOE Unconventional Gas Recovery Symposium of the Society of Petroleum Engineers, Pittsburgh, PA, 16-18 May. SPE/DOE-10847.

[8] Warpinski, N. R. and Teufel, L. W. 1987. Influence of Geologic Discontinuities on Hydraulic Fracture Propagation. J Pet Technol 39 (02).

http://dx.doi.org/10.2118/13224-PA [9] Shimizu, H., Hiyama, M., and Ito, T. 2014. Flow-coupled DEM Simulation for Hydraulic

Fracturing in Pre-Fractured Rock. Presented at the 48th US Rock Mechanics/Geomechanics Symposium, Minneapolis, 1-4 June. ARMA 14-7365.

[10] Pirayehgar, A. and Dusseault, M. B. 2014. The Stress Ratio Effect on Hydraulic Fracturing in The Presence of Natural Fractures. Presented at the 48th US Rock Mechanics/Geomechanics Symposium, Minneapolis, 1-4 June. ARMA 14-7138.

[11] Barenblatt, G. I., Zheltov, I. P., and Kochina, I. N. 1960. Basic Concepts in the Theory of Seepage of Homogeneous Liquids in Fissured Rocks (Strata). J Appl Math Mech 24 (5): 1286-1303.

http://dx.doi.org/10.1016/0021-8928(60)90107-6 [12] Kim, T. H. and Schechter, D. S. 2009. Estimation of Fracture Porosity of Naturally fractured

Reservoirs with No Matrix Porosity Using Fractal Discrete Fracture Networks. Presented at the 2007 SPE Annual Technical Conference and Exhibition, Anaheim, California, 11-14 November. SPE-110720.

101

Page 119: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[13] Yu, G. and Aguilera, R. 2012. 3D Analytical Modeling of Hydraulic Fracturing Stimulated rock volume. Presented at the SPE Latin American and Caribbean Petroleum Engineering Conference, Mexico City, Mexico, 16-18 April. SPE-153486.

[14] Izadi, G. and Elsworth, D. 2014. Reservoir stimulation and induced seismicity: Roles of fluid pressure and thermal transients on reactivated fractured networks. Geothermic 51: 368-379.

http://dx.doi.org/10.1016/j.geothermics.2014.01.014 [15] Valko, P. and Economides, M. J. 2001. Hydraulic Fracture Mechanics. West Sussex: John

Wiley and Sons. [16] Bratton, T. 2011. Hydraulic Fracture Complexity and Containment in Unconventional

Reservoirs. Oral presentation given at the ARMA Workshop on Rock Mechanics/Geomechanics, San Francisco, 23 June.

[17] Fang, Y., Elsworth, D., and Cladouhos, T. T. 2015, Estimating In-Situ Permeability of Stimulated EGS Reservoirs using MEQ Moment Magnitude: an Analysis of Newberry MEQ Data. Presented at the Fortieth Workshop on Geothermal Reservoir Engineering Stanford University, Stanford, California, 26-28 January.

[18] Core Laboratoies – Canada Ltd. 1985. Core Analysis of Amoco Canada Petroleum Company Ltd – Amoco et al WROSES 11-20-43-2W5 Glauconitic Formation in Westerose South Alberta. 16 October

[19] Jizba, D., Mavko, G., and Nur, M. 1990. Static and Dynamic Moduli of Tight Gas Sandstones. Paper was presented at SEG Conference, San Francisco, California, 23-27 September.

[20] Masters, J.A. 1970. Deep Basin gas trap, Western Canada. AAPG Bulletin 63 (2): 152-181. [21] Abaqus Release Notes 6.13, 2013, Dassault Systemes. [22] Nuth, M. and Laloui, L. 2007. Effective Stress Concept in Unsaturated Soils: Clarification

and Validation of a Unified Framework. Int J Numer Anal Meth Geomech 32 (7):771-801. http://dx.doi.org/10.1002/nag.645

[23] Maulianda, B.T., Hareland, G., and Chen, S. 2014. Geomechanical Consideration in Stimulated rock volume Dimension Models Prediction during Multi-Stage Hydraulic Fractures in Horizontal Wells – Glauconite Tight Formation in Hoadley Field. Presented at the 48th US Rock Mechanics/Geomechanics Symposium, Minneapolis, Minnesota, 1-4 June. ARMA 14-7449.

[24] Vermylen, J.P. 2011. Geomechanical Studies of the Barnett Shale, Texas, USA. Ph. D. Thesis. Stanford University, Stanford, California (May 2011).

[25] Brown, S.R. 1987. Fluid Flow through Rock Joints: The Effect of Surface Roughness. J of Geophys Res 92 (82):1337-1347. http://dx.doi.org/10.1029/JB092iB02p01337

102

Page 120: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

CHAPTER 4: GEOMECHANICAL CONSIDERATION IN

STIMULATED ROCK VOLUME DIMENSION MODELS

PREDICTION DURING MULTI-STAGE HYDRAULIC

FRACTURES IN HORIZONTAL WELLBORE –

GLAUCONITIC TIGHT FORMATION IN HOADLEY FIELD

Summary A fracture network or stimulated rock volume (SRV) within a tight rock reservoir can be induced

by hydraulic fracturing. The dimensions of the SRV induced by hydraulic fracturing are one of the

key measures of the success of the fracturing operation and set the volume of the reservoir

contacted for petroleum production. Thus, it is important to predict the dimensions of the SRV

given hydraulic fracture operating parameters to optimize recovery factor from the reservoir. In

this research, new analytical models are proposed to estimate SRV dimensions created from

hydraulically fractured horizontal well in unconventional reservoir. More specifically, the models

use the effective stresses, injected fluid volume, and other reservoir and hydraulic fracture injection

parameters to predict the dimensions of the SRV. Here, the model is calibrated by using

microseismic data from 6 stages of a hydraulic fracture job in a horizontal well penetrating the

Glauconitic Formation in the Hoadley Field, Alberta, Canada. The calibrated model for the

dimensions of the SRV can serve as an optimal fracture spacing estimator for future hydraulic

fracture job designs. The average estimated SRV width is smaller than the average fracture port

spacing and therefore for this study it is suggested to have the fracture port spacing tighter and

equal with the estimated SRV width for optimum design and future production.

103

Page 121: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.1 Introduction

BP Statistical review (2015) stated that the total world proved reserves of oil and natural gas

at the end of 2014 was 1.7 1012 barrels and 187.1 trillion cubic meters (tcm). The total world

production of oil and gas in 2014 was 88.7 106 barrels of oil per day (BOPD) (increased 2.3% from

2013) and 3.461 tcm per day (grew 1.6% from 2013) and the total world consumption of oil and

gas in 2013 was 92.1 106 BOPD (increased 0.8% from 2013) and 3.393 tcm per day (increased

0.4% from 2013) [1]. In 2014, BP Statistical review (2015) mentioned that it was about 91% of

the world’s daily energy consumption was still derived from oil, gas, coal, and nuclear power with

32.6% from oil and 23.7% from gas [1]. The increase of the consumption of oil and gas together

with the decline of producing fields requires that new resources (unconventional reservoirs) must

be developed. In North America, much of this demand of oil and gas will be filled from production

of tight oil and tight gas resources.

Unconventional reservoirs are oil and gas deposits that cannot be produced at economic flow

rates or do not produce at economic volumes without an assistance from massive stimulation

treatments such as hydraulic fracture or steam injection. Unconventional tight rock reservoirs are

the focus of our study, more specifically to optimize hydraulic fracture design in horizontal wells

in tight gas reservoirs.

Hydraulic fracturing is used to increase the effective drainage area and the permeability of the

reservoir by creating a fracture network around the well often referred to as a stimulated rock

volume (SRV). The dimensions of the SRV are one of the main controls of horizontal well

performance after the hydraulic fracture operation has been performed since it sets the volume of

the enhanced permeability zone within the reservoir and provides an estimate of the stimulated

region of the reservoir which in turn sets the potential volume of petroleum fluids that can be

produced. It is important to determine the SRV dimensions to identify possible optimum hydraulic

fracture treatment parameters, for example fracture port spacing and injected fluid volumes.

104

Page 122: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.1.1 Objective of Study

In this study, a new analytical model to determine the dimensions of the SRV is derived which

integrates reservoir rock mechanical properties such as Young’s modulus, Poisson’s ratio, internal

friction angle, cohesion, effective fractured zone permeability and porosity, effective stresses, and

injected fluid volume during the hydraulic fracture operation.

4.2 Literature Review

4.2.1 Failure Mechanics

Rock failure occurs when a large stress is applied to a rock leading to a permanent change of

shape of the rock and its integrity. The failure state is often accompanied with much lower

capability to carry loads [2]. The stress level at which the rock failed is called the rock strength

and it is usually determined in the laboratory using uniaxial or triaxial tests.

4.2.1.1 Tensile Failure

Tensile failure happens when the effective stress across some plane within the rock exceeds a

critical limit referred to as the tensile strength [2]. Lockner (1995) mentioned that the tensile

strength for most rocks were low (of order of a few MPa), and when there were natural fractures

in the rock, the tensile strength, T0, was expected to be near zero [3]. The minimum effective stress,

σ3’, is given by (Fjaer et al. 2008) [2]:

𝜎𝜎3′ = −𝑇𝑇0 (1)

Fjaer et al. 2008 stated that a hydraulic fracture was a form of tensile failure that occurred when

the fluid pressure exceeded the sum of the minimum total stress and the tensile strength of the rock

[2]. Zoback (2007) stated that tensile failure extension occurred when the injection pressure was

higher than the minimum stress [4]. Continuous pumping of fluid into the rock at high pressure

causes the fracture to grow in the direction of least resistance within the rock; this is the direction

normal to the minimum stress in the rock.

105

Page 123: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.2.1.2 Shear Failure

Fjaer et al. (2008) explained that shear failure happened if the shear stress along some planes

in the rock was high enough, and developed a failed zone along the failure plane where the two

sides of the plane moved relatively to each other by friction as shown in (examples are illustrated

in Figure 2.1b in Chapter 2) [2].

Shear failure can be determined by using a Mohr-Coulomb failure envelope. Jaeger and Cook

(1979) explained that the failure envelope was built from the cohesion as the intercept and the

internal friction angle as the slope [5]. High cohesion and internal friction angle are typical of

strong rocks which are hard to fail. A naturally fractured reservoir is a relatively weak rock with

lower cohesion which is easier to fail than strong rocks. The Mohr circle consists of maximum and

minimum effective stresses. Zoback (2007) explained that the Mohr-Coulomb failure envelope

was produced by using test results from triaxial tests (illustrated in Figures 2.2a and 2.2b in Chapter

2) [4]. Triaxial tests involve applying a load on the sample (σ1) while the confining pressure (σ3)

is held constant until the sample fails. The Mohr-Coulomb failure envelope slope usually decreases

for most rocks as the confining pressure increases. But for most rocks, it is allowable to consider

a linearized Mohr-Coulomb failure envelope (illustrated in Figure 2.2c in Chapter 2) [4]. The

linearized Mohr-Coulomb failure envelope criterion is [4]:

𝜏𝜏 = 𝐶𝐶 + 𝜎𝜎𝑛𝑛𝜇𝜇𝑖𝑖 (2)

𝜏𝜏 = 𝐶𝐶 + 𝜎𝜎𝑛𝑛𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 (3)

where τ is the shear stress, C is the rock cohesion, σn is the normal stress, µi is the slope of the

failure envelope, ϕ is the internal friction angle. The normal stress on a failure plane is inclined at

an angle β to the least stress 𝜎𝜎3, where the minor principal stress is 𝜎𝜎3 (illustrated in Figure 2.2 in

Chapter 2) [2, 4]:

𝜎𝜎𝑛𝑛 = 𝜎𝜎1+𝜎𝜎32

+ 𝜎𝜎1−𝜎𝜎32

𝑐𝑐𝑐𝑐𝑐𝑐2𝛽𝛽 (4)

𝜏𝜏 = 𝜎𝜎1−𝜎𝜎32

𝑐𝑐𝑠𝑠𝑡𝑡2𝛽𝛽 (5)

2𝛽𝛽 = 90𝑜𝑜 + 𝑡𝑡 (6)

106

Page 124: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

𝛽𝛽 = 45𝑜𝑜 + 𝜙𝜙2 (7)

Shear failure of rocks during hydraulic fracturing is induced by the increase of pore pressure.

Warpinski and Branagan (1988) presented that the shear failure occurred during hydraulic

fracturing at some distance from the center of the main hydraulic fracture [6]. Warpinski and

Branagan (1988) used geomechanics to predict the shear failure and revealed that the extent of the

shear failure zone was affected by the fracture pressure. Their results were verified by using

microseismic monitoring data.

Jupe et al. (1993) showed, from the microseismic data of a geothermal site showed that the

dominant mechanism that enhanced SRV permeability was shear failure along pre-existing natural

fractures [7].

Duchane (1998) showed, by using the microseismic observation and geological information of

a geothermal site, that the hydraulic vertical fractures were created from reopened sealed natural

fractures (tensile and shear failures) instead of inducing new hydraulic vertical fractures (tensile

fracture) [8].

Rahman et al. (2002) developed a model that combined tensile-induced hydraulic fractures and

shear failure of natural fractures [9]. Rahman et al. (2002) revealed that the enhanced permeability

of the SRV was 30 times that of the initial formation permeability.

Palmer et al. (2007) discovered shear failures at natural fractures far from the central hydraulic

fracture during hydraulic fracturing in the Barnett shale [10]. They also determined that the

enhanced permeability induced from hydraulic fracturing and the permeability during production

might not be equal because the shear or tensile fractures that were induced during the operation

might partially close during production [10].

4.2.2 Stimulated Rock Volume (SRV) Models for Tight Rock

Unconventional Reservoirs

To design hydraulic fracture jobs, it is necessary to predict the growth of the SRV as a function

of hydraulic fracture injection parameters. SRV growth in heterogeneous formations with low

stress anisotropy is complex and it remains difficult to be predicted with a high level of certainty.

107

Page 125: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

In this study of the Hoadley Field, it is observed that the difference between the maximum and

intermediate stresses is small whereas the difference between the maximum and minimum stresses

is large.

Throughout the years, various models have been developed to approximate fracture geometry

during hydraulic fracturing operations. Fracture models can be separated into two-dimensional

(2D) and three-dimensional (3D) categories. The KGD (Khristianovitch and Zheltov 1955;

Geertsma and de Klerk 1969) [11, 12] and the PKN (Perkins and Kern 1961) models [13]

(illustrated in Figure 2.7 in Chapter 2) are the most popular 2D models as explained by Rahman

and Rahman (2010) [14]. Barree (2009) stated that in all 2D models, only the fracture width and

length were derived from the models while the fracture height remained constant [15]. The PKN

and KGD models both use the Sneddon (1946) solution [16]. Sneddon (1946) proposed an equation

for the width of a crack [16]:

𝑤𝑤 = 2𝑢𝑢 = 4�1−𝑣𝑣2�𝑃𝑃0𝜋𝜋𝜋𝜋

𝑐𝑐 (8)

where w is the crack width, u is the crack half width, v is the Poisson’s ratio, P is the applied

pressure, E is the Young’s modulus, and c is the crack half length. Nordgren (1970), Barree (2009),

and Rahman and Rahman (2010) summarized the different assumptions for the PKN and KGD

models were summarized [14, 15, 17]. The PKN model assumptions were:

5. The crack length was larger than the crack height.

6. The crack height was restricted to a limited section due to the presence of upper and lower

barriers.

7. There was no vertical extension in each vertical section; therefore the fracture shape was

elliptical.

8. A 2D plane strain deformation was created in the vertical plane.

The PKN equation provided an estimate of the crack width as proposed by Perkins and Kern (1961)

[13]:

𝑤𝑤 = 2𝑢𝑢 = 2�1−𝑣𝑣2��𝑃𝑃𝑓𝑓−𝜎𝜎3�𝜋𝜋

𝐻𝐻 (9)

where H is the crack height, and Pf is the fracture fluid pressure.

The KGD model assumptions were:

108

Page 126: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5. The crack height is larger than the crack length.

6. The crack height is constant and uniform along the crack length; therefore the fracture

shape is rectangular.

7. The crack width is constant in the vertical direction.

8. A 2D plane strain deformation is created in the horizontal plane.

The KGD equation produced the crack width as mentioned by Daneshy (1971) [18]:

𝑤𝑤 = 2𝑢𝑢 = 2�1−𝑣𝑣2��𝑃𝑃𝑓𝑓−𝜎𝜎3�𝜋𝜋

𝑥𝑥𝑓𝑓 (10)

where xf is the crack half-length.

Economides and Nolte (2000) summarized the available models such as planar 3D, pseudo 3D,

and general 3D models [19]. The planar 3D model assumed that the fracture was planar and

perpendicular to the minimum stress. This model was applicable when the surrounding zones had

stresses lower or similar to the stresses of the interest formation. The pseudo 3D model was divided

into lumped and cell based models. The lumped based model had the two half-ellipsis join at the

center in the vertical profile [19]. The fracture half-length and height were calculated at each time

step with the assumed shape was elliptical. The cell based model considered the fractures as a

series of connected cells. These models did not have fixed shapes but they assumed a plane strain

model and they did not fully couple fluid flow in vertical direction to fracture geometry. The

general 3D model did not have any assumption on the fracture orientation. The fracture orientation

was determined by the wellbore perforations and orientations and state of stress.

There were several analytical pseudo 3D and general 3D models proposed recently. Fisher et

al. (2002) conducted the hydraulic fracture diagnostic projects in the naturally fractured Barnett

shale reservoirs [20]. They showed that the fracture half-length depended on the injected fluid

volume where the fracture half-length stopped growing after a significant amount of injected fluid

volume was injected. The SRV half-length and width were observed by using microseismic

monitoring.

Maxwell et al. (2002) from the microseismic monitoring during the hydraulic fracturing in the

Barnett shale, discovered that hydraulic fracturing occasionally grew at an angle to the assumed

fracture direction (the maximum stress direction) and into neighboring wells [21]. The results

showed that the hydraulic fractures grew at an angle because they intersected the natural fractures

109

Page 127: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

network. They also discovered that the hydraulic fracture grew to the neighboring wells because

the depleted zones around the neighboring wells.

Xu et al. (2009) used a semi-analytical pseudo 3D geomechanical model to study the

interaction between fractures and injected fluid volume by presenting the hydraulic fracture as a

horizontally expanding ellipse fractures [22]. They found that the fracture network complexity and

its dimensions were affected by the ratio of stresses within the reservoir.

Maxwell et al. (2010) showed that there were some cases where critically stressed fractures

close to the point of hydraulic fracture deformation could trigger small stress changes that result

in remote trigger of microseismic events [23]. This explained one of the causes of microseismic

measurement uncertainties and an overestimated SRV.

Mayerhofer et al. (2010) predicted the extent of the stimulated rock area (SRA) after hydraulic

fracturing by using microseismic mapping in the horizontal well [24]. They drew a constant width

rectangle in the direction of maximum stress from the wellbore to the farthest event in the rectangle

on both sides of horizontal well, and then they estimated the SRV height for the individual

container. The limitation of this method was the requirement of adequate microseismic events and

it was only applied to a particular field.

Weng et al. (2011) developed an analytical 3D fracture network numerical model in the

naturally fractured reservoir to determine the dimensions of the SRV [25]. Their simulation results

showed that stress anisotropy, natural fractures, and internal friction angle affected the complexity

of the fracture network. By lowering the stress anisotropy, the fracture changed from a bi-wing

fracture into a complex fracture network.

Yu and Aguilera (2012) presented an analytical 3D model to determine the dimensions of a

SRV after a hydraulic fracturing operation in an unconventional gas reservoir by using

microseismic events and the pressure diffusivity equation [26]. Their results provided the SRV

dimensions as a function of injection pressure, minimum pressure to trigger microseismic events,

microseismic event occurrence time, and hydraulic diffusivity coefficient. They determined the

hydraulic diffusivity coefficient required to calibrate the model and then predicted the dimensions

of the SRV. The hydraulic diffusivity coefficient could be determined from a slope of a straight

line plot between the distances between the microseismic events and the wellbore versus the square

root of the occurrence time of the microseismic event.

110

Page 128: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Nassir et al. (2012) developed a 3D geomechanical finite element model (FEM) to examine

SRV propagation in tight formations [27]. They found that their shapes of the SRV were similar

to that obtained from microseismic monitoring. They explained that the SRV dimensions

complexity was dependent on low rock cohesion and high initial contrast between the minimum

and maximum stresses. Their simulation results suggested that a rock with a low rock cohesion

(less than 1 MPa) produced wider SRVs. The main conclusion was that large and wide SRVs

would only occur in the formations where the rock was weakened by natural fractures (low rock

cohesion).

McClure and Horne (2013) developed a computational model that coupled fluid flow, stresses

and deformation induced by fracture opening/sliding, and fracture propagation in a 2D discrete

fracture network [28]. The model was able to couple fluid flow and earthquake models. The model

was used to investigate the interaction between the fluid flow, the permeability evolution, and the

induced seismicity during the hydraulic fracture injection into a single fault. Using this model,

they explained the critical importance of including stresses induced by the deformation in the

hydraulic fracture modeling. These stresses directly impacted the mechanism of the hydraulic

fracture propagation and the resulting fracture network properties. The key limitations of the model

was that it was a 2D model and it required the paths of newly forming fractures to be specified in

advance.

4.2.3 Microseismic Monitoring during Hydraulic Fracturing

A microseismic event is a micro-earthquake that happens during hydraulic fracturing. The

precise location of the microseismic event is defined as the location of a new fracture or an existing

fracture when it is reopened. The time at which the microseismic event is detected at the receiver

is the time which the P (compression) and S (shear) waves travel the distance from the event

location to the receiver’s location. The wave velocity models for different formations are built by

using a dipole sonic log and a perforation shot arrival time. The microseismic event location is

determined by using the distance between the sensor and the microseismic event based on the P

and S wave picks and also using the orientation (azimuth and dip) determined from a polarization

analysis (generated particle trajectory from wave propagation that is characterized by direction and

shape) [29].

111

Page 129: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Microseismic monitoring during hydraulic fracturing is passive measurement of microseismic

events and it provides microseismic event arrival time, location, and magnitude. The growth of the

dimensions of the SRV during hydraulic fracturing is a key criteria of the effectiveness of hydraulic

fracturing. The SRV dimensions can be estimated by using the microseismic events locations.

The first microseismic monitoring application on the hydraulic fracture was the Rangely

experiments [30]. The experiments were conducted on controlled fluid injection to detect the

induced microseismicity. The downhole and surface microseismic monitoring were used. The

downhole monitoring used arrays of geophones in a nearby observation wellbore and the surface

monitoring used surface array sensors.

Downhole microseismic monitoring is the main direct observation method to monitor the

hydraulic fracture dimensions at depth. Two techniques are available to detect microseismic events

downhole which are downhole receiver array and downhole tiltmeter array. Though these

techniques have strength and limitations but they provide valuable information under certain

conditions.

Downhole monitoring using receiver array procedures are place an array of triaxial geophone

or accelerometer receiver downhole in an observation well in a certain orientation, record

microseismic data during hydraulic fracture, determine the microseismic events within the data

catalog, locate those events, and interpret the whole microseismic data catalog[31, 32, 33, 34].

Wright (1998) explained that tiltmeters identified changes in the sensor’s angular position [35].

Warpinski et al. (2006) mentioned that the sensor was very sensitive with a sensitivity equivalent

to 0.2 inch movement over a 3,000 mile range [36]. The angular position provided a measure of

the earth deformation process [36]. The sensor only measured the tilt along one axis therefore it

was required to have two orthogonal sensors to provide a full tilt measurements (magnitude and

angle) [36]. Warpinski et al. (2006) explained that the hydraulic fracturing produced tilt signatures

that were inverted to define the dimensions of the fractures [36].

Surface microseismic monitoring is done by placing a large number of arrays on the surface.

Hall and Kilpatrick (2009) conducted experiments by using surface arrays consisting of 1,078

stations of 12 geophones spread out in a radial pattern around the hydraulic fracture well [37].

These geophones were buried to a depth of one foot to get maximum signal-to-noise ratio by

reducing rainfall interference (illustrated in Figure 2.11 in Chapter 2) [38].

112

Page 130: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.3 Hoadley Field Project

ConocoPhillips Canada conducted twelve stages of hydraulic fracturing in two horizontal wells

near Rimbey, Alberta, Canada as shown in Figure 4.1 [39]. Both wells, Wells 1 and 2, were

monitored by using microseismic with a 12 sensor vertical array in a nearby vertical well (Figure

4.2). This study focuses on Well 1-18-43-2W5.

The formation of interest was the Glauconitic Formation with a thickness of 43 m. Underneath

was the Ostracod with a thickness of 10 m. The formations above the Glauconitic was the Medicine

River Coal with a thickness of 5 m and the Mannville with a thickness of 80 m. The horizontal

well was drilled in the Glauconitic Formation at 1,900 mTVD with a lateral section of 2,000 mMD

[39]. There were a total of 1,660 events in the final processing dataset (Figure 4.3). The total events

used to determine the dimensions of the SRV from Well 1 were 732 events (Figure 4.4). The

majorities of the hydraulic fractures propagated upward starting from the horizontal wellbore and

grew to the Medicine River Coal and the Mannville formations with several hydraulic fractures

propagating downward to the Ostracod formation. The fracture dimensions were only interpreted

via microseismic data for 6 stages on Well 1-18-43-2W5.

113

Page 131: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 4.1: (a) The site location and the maximum stress direction (45o NE) and (b) hydraulic fracture treatment location near Red Deer [39].

HZ 1-18-43-2W5

10-18-43-2W5

6-18-43-2W5

HZ 4-18-43-2W5

114

Page 132: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.2: Microseismic downhole monitoring array configuration in nearby wellbore [39].

Fig. 4.3: Depth distribution of microseismic events from two horizontal wellbores hydraulic fracture [39].

115

Page 133: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.4: Executive summary of the two treatment wellbores, observation wellbore, producing wellbore and observed microseismic events during 12 stages of hydraulic fracture [39].

The determination of the Hoadley field stress regime is based on the world stress map. Most

of the North Western part of the North America are under compression which causes high

horizontal stress. Therefore, the greatest stress in Figure 4.1 is generally perpendicular to the front

ranges in a SW-NE direction and it is often the maximum horizontal stress. In Alberta the least

principal stress is often horizontal [40, 41]. Therefore, the Hoadley field is assumed to be in a

strike slip fault regime with the greatest stress being the maximum horizontal stress and the least

stress being the minimum horizontal stress.

The regional stress map shows that the maximum stress direction is 45°NE (± 3°) (Figure 4.1a).

This direction is parallel to the direction of the observed hydraulic fracture (48°NE with ± 3°)

based on the microseismic events. This indicates that the direction of fracture growth is mainly

controlled by the maximum horizontal stress direction.

116

Page 134: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The data used in this study are as follows: formation properties, open hole logs, hydraulic

fracture injection report, interpreted microseismic events locations, horizontal well trajectory, and

interpreted fracture dimensions via microseismic from 6 hydraulic fracture stages. Figure 4.5

shows the bottomhole pressures (BHP) from the hydraulic fracture injection report [39].

Fig. 4.5: Bottomhole injection pressure for horizontal wellbore A during 12 stages of hydraulic fracture [39].

Each hydraulic fracture injection stage comprised pad, slurry, spacer and flush injection stages.

The pad stage consisted of linked gel (WF130) and gas (N2), the slurry stage consisted of linked

gel (WF130) and gas (N2) with proppants (Jordan unimen of 20/40 mesh with median particle

diameter of 0.662 mm and Jordan unimen of 100 mesh with median particle diameter of 0.19 mm),

the spacer stage consisted of linked gel (WF130) and gas (N2), and the flush stage consisted of

slick water. These fluid properties are listed on Table 4.1. The total average slurry injection volume

rates for the 12 stages was 5 m3/min.

20

25

30

35

40

45

50

55

60

65

0:00:00 0:07:12 0:14:24 0:21:36 0:28:48 0:36:00 0:43:12 0:50:24 0:57:36

BHP (MPa)

Time (min:sec)

Stage 1Stage 2Stage 3Stage 4Stage 5Stage 6Stage 7Stage 8Stage 9Stage 10Stage 11Stage 12

117

Page 135: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 4.1: Formation and hydraulic fracture fluid properties [39].

Injection flow rate (bpd) 45286.65 Fluid viscosity (cP) 30 Horizontal wellbore radius (ft) 0.08 Stimulation skin -7.00 Proppant type Jordan Unimen Proppant mesh size 20/40, 100 Hydraulic fracture fluid WF130 slickwater Gas Nitrogen

4.4 Research Workflow

4.4.1 Derivation of Equations for the SRV Dimensions

The model derived here is based on a SRV in a reservoir with a low rock cohesion and a high

initial contrast between the stresses by Nassir et al. (2012) [27]. Nassir et al. (2012) results

suggested that a rock with a low rock cohesion (less than 1 MPa) produced wide SRV [27].

Therefore, their conclusions were that large and wide SRV would only be found in the formations

where the rock was weakened by natural fractures (low rock cohesion).

Wider SRV was caused by the closeness of the initial reservoir conditions to the shear failure

criterion (naturally fractured reservoir with low cohesion) and high initial stress contrast. The

tendency to cause shear failure led to a more complex fracture network rather than a simple bi-

wing hydraulic fracture.

According to Rich and Ammerman (2010) and Bratton (2011), fracture geometry complexity

was also to be found as a function of the state of stress [42, 43]. When reducing the anisotropy, a

more complex fracture network was created because there was no preferred direction. In this study,

the maximum horizontal stress and the overburden stress ratio was small (1.07) which led to

fractures orientations in the maximum horizontal stress and vertical directions.

Perkins and Kern (1961) presented an estimate of a crack width (fracture aperture) model as a

function of the injected flow rate, the plane strain modulus, and the minimum effective stress [13].

Integrating the previous work by Perkins and Kern (1961), Rich and Ammerman (2010), Bratton

(2011), and Nassir et al. (2012), [13, 27, 42, 43], the proposed dimensions of the SRV under tensile

118

Page 136: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

failure are derived as a function of effective stress, injected fluid volume and plane strain modulus

as follows:

𝑡𝑡 = 𝑘𝑘1 �𝑉𝑉𝑖𝑖𝜋𝜋′𝜎𝜎𝐻𝐻′ �

13 (11)

𝑛𝑛 = 𝑘𝑘2 �𝑉𝑉𝑖𝑖𝜋𝜋′𝜎𝜎ℎ′ �

13 (12)

𝑐𝑐 = 𝑘𝑘3 �𝑉𝑉𝑖𝑖𝜋𝜋′𝜎𝜎𝑣𝑣′�

13 (13)

where a, b, and c are the SRV length, width, and height, k1, k2 and k3 are assumed isotropic

constants from the microseismic calibration, Vi is the injected volume, and E’ is the plane strain

modulus.

The proposed SRV dimensions under shear failures are derived by combining the previous

work by Perkins and Kern (1961), Rich and Ammerman (2010), Bratton (2011) [13, 42, 43], and

the shear failure criterion (the effective stresses have linear relationship with the rock cohesion

and the internal friction angle) (illustrated in triangle OAB in Figure 2.4 in Chapter 2).

𝑐𝑐𝑐𝑐𝑐𝑐𝑡𝑡 =12

(𝜎𝜎1−𝜎𝜎3)

�𝐶𝐶+12(𝜎𝜎1+𝜎𝜎3)𝑡𝑡𝑡𝑡𝑛𝑛𝜙𝜙� (14)

Rearranging Equation (14) yields:

12

(𝜎𝜎1 − 𝜎𝜎3) = 𝐶𝐶𝑐𝑐𝑐𝑐𝑐𝑐𝑡𝑡 + 12

(𝜎𝜎1 + 𝜎𝜎3)𝑐𝑐𝑠𝑠𝑡𝑡𝑡𝑡 (15)

𝜎𝜎1 = 2𝐶𝐶𝐶𝐶𝑜𝑜𝐶𝐶𝜙𝜙(1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙) + 𝜎𝜎3(1+𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙)

(1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙) (16)

Thus, the dimensions of the SRV during the shear failure becomes:

𝑡𝑡 = 𝑘𝑘1𝜎𝜎𝐻𝐻′(𝑉𝑉𝑖𝑖)

13

�2𝐶𝐶𝐶𝐶𝑜𝑜𝐶𝐶𝐶𝐶1−𝐶𝐶𝑖𝑖𝑠𝑠𝐶𝐶� (17)

𝑛𝑛 = 𝑘𝑘2𝜎𝜎ℎ′(𝑉𝑉𝑖𝑖)

13

�2𝐶𝐶𝐶𝐶𝑜𝑜𝐶𝐶𝐶𝐶1−𝐶𝐶𝑖𝑖𝑠𝑠𝐶𝐶� (18)

𝑐𝑐 = 𝑘𝑘3𝜎𝜎𝑣𝑣′(𝑉𝑉𝑖𝑖)

13

�2𝐶𝐶𝐶𝐶𝑜𝑜𝐶𝐶𝐶𝐶1−𝐶𝐶𝑖𝑖𝑠𝑠𝐶𝐶� (19)

119

Page 137: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

where C is the naturally fractured reservoir cohesion which is lower than the intact rock cohesion,

and 𝑡𝑡 is the naturally fractured reservoir internal friction angle.

4.4.2 Analysis of Microseismic Events

The next step is to divide the microseismic events into n time steps. This workflow assumes

there is no difference between the time at which the hydraulic fracture injection times are recorded

and the time at which the microseismic events are recorded. The hydraulic fracture injection time

and the corresponding microseismic events for each hydraulic fracture stage are divided into two

time steps as follows:

1. A first time step with injection time and microseismic events that represent tensile failure

with a high pressure (fracture breakdown pressure) and a short injection time. The tensile

failure pressure and injection time were measured in the field. The tensile failure was

assumed to occur in the proximity of the horizontal wellbore with no pressure drop from

the horizontal wellbore to the tensile failure location. The pressure drop was neglected due

to low flow resistance in the highly conductive fractures and associated SRV.

2. A second time step with injection time and microseismic events that represent shear failure

with a lower pressure. The low pressure is assumed to meet the shear failure criterion. The

shear failure pressure experiences pressure drop from the horizontal wellbore to shear

failure location (natural fracture location). The shear failure pressure is estimated from

Chapter 3. The shear failure is assumed to occur from the start of fracture propagation

pressure until the end of injection.

In this procedure, growth of the SRV is based on the movement of the microseismic event

center from the first time step to the second time step. Each time step microseismic event center is

estimated by using the average of the event locations from previous time step.

There are only six hydraulic fracture stages with the microseismic interpreted SRV dimensions

available from ConocoPhillips Canada. These six stages were analyzed to determine which stage

is the best representative for the other eleven stages. Stage 7 is determined as the best

representative stage because:

120

Page 138: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

1. There are adequate microseismic events (4 events) during the first time step in the

proximity of the horizontal wellbore to be estimated as SRV dimensions during the first

time step.

2. There are adequate microseismic events (41 events) during the second time step to be

estimated as SRV dimensions during the second time step.

3. First time step events locations are in the proximity of the horizontal wellbore that shows

an apparent SRV growth from the horizontal wellbore to the first and second time step.

4. Stage 7 is located far from an old depleted producing well on the North East side of the

horizontal wellbore. This well has impacted stages 8 and 9 by triggering more events and

creating bigger SRV compare to estimated SRV.

5. Stage 7 is located not far from the microseismic observation well on the West side of the

horizontal wellbore, but this well has not affected Stage 7.

6. Stage 7 fracture breakdown BHP of 41.5 MPa is close to the average fracture breakdown

BHP of 46.65 MPa interpreted for six stages.

7. Stage 7 microseismic interpreted SRV length and width ratio of 2.9 is close to the average

for twelve stages microseismic maximum and minimum stresses ratio of 3.

A twenty metre cutoff is used to consider that the first time step events are in the proximity of

the horizontal wellbore. The cutoff is estimated from the average first time step events of

microseismic interpreted stages (Stages 6, 7, 9, 10, 11, and 12). If the first time step events are

within the cutoff, then the SRV growth is considered to be apparent.

Most of the microseismic events occur on the East side (48°NE) of the horizontal wellbore

which is 3° different from the maximum horizontal stress direction (45°NE). This finding is

consistent with Maxwell et al. (2002) [21]. They explained that the hydraulic fracture orientation

could be different from the maximum stress orientation. This phenomenon could be caused by the

different property values in the maximum and the minimum horizontal stress directions. The

property is the diffusivity coefficient. The diffusivity coefficient is a function of equivalent

permeability, equivalent compressibility, equivalent Young’s modulus, and equivalent Poisson’s

ratio. These properties are estimated by using the proposed equations of the diffusivity ratio in the

maximum and minimum horizontal stress directions. Yu and Aguilera (2012) also proposed the

SRV dimensions as a function of the hydraulic diffusivity [26].

121

Page 139: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.4.3 Input Parameters

The input parameters used to estimate the dimensions of the SRV include the total injected

fluid volume, BHP, Poisson’s ratio, Young’s modulus, cohesion, internal friction angle, and

effective stresses. The Poisson’s ratio and the Young’s modulus are calculated by using the

compressional and the shear wave velocities from logs. The Young’s modulus and the Poisson’s

ratio are required to calculate the plane strain modulus. The plane strain modulus is an input for

the tensile failure SRV dimensions model.

𝐸𝐸′ = 𝜋𝜋1−𝑣𝑣2

(20)

Where E’ is plane strain modulus, E is Young’s modulus, and v is Poisson’s ratio. The naturally

fractured reservoir cohesion and internal friction angle are needed to calculate the shear failure

SRV dimensions. The naturally fractured reservoir cohesion is determined from the intercept of

Mohr-Coulomb failure envelope when the Mohr circle touches the envelope.

The pore pressure at shear failure is predicted from Chapter 3 because there is no field data

available. In this chapter, the field BHP data is used to determine:

1. Tensile failure pressure (fracture breakdown pressure) and duration.

2. Shear failure duration is from 5.5 minutes after hydraulic fracture starts (starts of hydraulic

fracture propagation) until the end of injection.

The shear failure pressure of 12.12 MPa is estimated at 5.5 minutes after hydraulic fracture

starts and distance of 6.55 m (estimated half natural fracture distance) from Chapter 3 as shown in

Figure 4.6a. The horizontal wellbore is assumed to be in the middle of two natural fractures.

The internal friction angle is calculated using the natural fracture angle equation (it is explained

in the next section). The Mohr circle is built using the maximum and minimum effective stresses

and the Mohr circle radius.

The minimum horizontal stress gradient was determined for Glauconitic Formation with a

value of 11.66 kPa/m (0.52 psi/ft) from the instantaneous shut-in pressure [39]. This value was

similar to values from nearby wells [42]. The overburden stress gradient of 24.09 kPa/m (1.07

psi/ft) was estimated using the density log from the formation depth of 1,900 m until 1m923.5 m.

The maximum horizontal stress gradient of 25.79 kPa/m (1.14 psi/ft) was determined by using a

122

Page 140: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

commercial geomechanics software analysis package (STABView). The inputs were the

minimum and maximum horizontal and the overburden stress gradients, the initial formation pore

pressure gradient, and the horizontal wellbore radius. The simulation output was the wellbore

breakout (from caliper logs) at the nearby vertical wellbore. The maximum horizontal stress was

determined by simulating the inputs that produced the wellbore breakout matched the nearby

vertical wellbore breakout. The initial pore pressure gradient of 4.86 kPa/m (0.21 psi/ft) was

determined from the pressure build up test in horizontal Well 1 [39].

The microseismic events showed that the natural fracture plane was at 30° inclined (θ) with

respect to the maximum horizontal stress. This observation led to the calculation of normal stress

at natural fracture plane inclination with respect to maximum horizontal stress of 60° (β):

𝛽𝛽 = 90𝑜𝑜 − 𝜃𝜃 (20)

The internal friction angle of the naturally fractured reservoir rock was assumed to be equal

with the intact rock internal friction angle (ϕ). The internal friction angle of 30° was calculated by

using Equation (7). This study assumed the natural fracture was reopened by the shear failure

induced by the hydraulic fracture.

The sonic travel time of 263.12 μs/m (80.2 μs/ft) was used to determine the intact rock

unconfined compressive strength (UCS) of 40 MPa using empirical relationship from Hareland

and Nygaard (2007) [44]. The internal friction angle and the UCS were used to calculate the intact

rock cohesion (C) of 11.54 MPa using the equation from (Goodarzi and Settari 2009) [45]:

𝐶𝐶 = 𝑈𝑈𝐶𝐶𝑆𝑆(1−𝐶𝐶𝑖𝑖𝑛𝑛𝜙𝜙)2𝐶𝐶𝑜𝑜𝐶𝐶𝜙𝜙

(21)

The calculated intact rock cohesion was too high for the Mohr circle to touch the Mohr-Coulomb

failure envelope and create shear failures. The natural fractures caused a lower rock cohesion.

Hoek (1983) explained that the natural fracture lowered the formation shear strength and the

formation strength was defined by the naturally fractured reservoir internal friction angle and

cohesion [46]. The lowest strength occurred when the natural fracture was at a certain inclination

with respect to the maximum stress. To support this, Hoek (1983) provided a set of triaxial tests

results for fractured sandstone with the lowest strength occurred when the natural fracture plane

was at 30° inclined with respect to the maximum horizontal stress [46].

123

Page 141: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The cohesion of the naturally fractured reservoir of 0.69 MPa (100 psi) was estimated from the

intercept of Mohr-Coulomb failure envelope when the Mohr circle touched the failure envelope.

The Mohr circle is determined using the shear failure pore pressure from Chapter 3, maximum

horizontal stress, and minimum horizontal stress. The complete inputs for the models simulation

are shown in Table 4.2.

Table 4.2: Glauconitic Formation properties.

Parameters Values Units Minimum horizontal stress gradient 11.66 kPa/m Maximum horizontal stress gradient 25.79 kPa/m Overburden stress gradient 24.09 kPa/m Biot’s coefficient 1 Average depth 1900 m Poisson’s ratio 0.23 Porosity 15 % Intact rock and naturally fractured rock friction angle 30 degree Intact rock cohesion 11.55 MPa

4.4.4 Calibration of Stimulated Rock Volume using Microseismic Data

The calibration constants are determined by finding the constants that match the estimated

Stage 7 SRV dimensions with the Stage 7 SRV dimensions interpreted from microseismic events.

The failure type needs to be determined prior to calibrating the model. The first time step BHP is

required to initiate the fracture and associated SRV via tensile failure. The first time step BHP and

hydraulic fracture injection duration are different for all stages that lead to different first time step

SRV dimensions. Stages 1, 11, and 12 have the highest tensile failure BHP compared to the other

stages; this might be caused by less intersection with the natural fractures compared to the other

stages.

The pore pressure at which tensile failure occurs in the first time step of Stage 7 is equal to

41.5 MPa. The pore pressure at which shear failure occurs in the second time step of Stage 7 is

12.12 MPa as shown in Figure 4.6a (330 s and 6.55 m). The Mohr circle for original reservoir

pressure assumes a Biot’s constant of unity and no shear failure. In this study, a linear

simplification of the Mohr-Coulomb failure envelope is used as shown in Figure 4.6b [4]. The

124

Page 142: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

calibration constants are k1, k2 and k3. The calibration constants determined for the first time step

under tensile failure are k1 = 200, k2 = 100 and k3 = 150. The calibration constants found for the

second time step under shear failure are k1 = 0.6485, k2 = 0.3336, and k3 = 0.1474. These constants

are applicable for the Glauconitic Formation at Hoadley field or other formations that have similar

reservoir and geomechanical properties.

125

Page 143: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 4.6: (a) Pressure drop derived from finite element analysis in Chapter 3 and (b) Stage 7 Mohr-Coulomb failure envelope for Case 1.

126

Page 144: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.5 Results

4.5.1 Constant Total Stress

The first results are produced using constant total stresses (Case 1). The SRV length, width

and height for Stage 7 are calibrated to match the closest first time step event. The second time

step SRV length, width and height for Stage 7 are calibrated to match SRV dimensions interpreted

via microseismic events. The Stage 7 calibration constants for the tensile and shear time steps (k1,

k2, and k3) are applied to the other stages.

Here, the distances obtained from the first time step events are compared to the cutoff, SRV

growth (yes or no), number of events in the estimated SRV, and number of events in the region

double the size of the estimated SRV. The comparison is presented in Table 4.3. The SRV

dimensions for the six stages with dimensions interpreted from microseismic observations are

shown in Figures 4.7 to 4.15. The SRV dimensions estimated with no interpretations from

microseismic observations are shown in Figures 4.16 to 4.21.

Table 4.3: Results of all stages first time step event distance, SRV growth, number of events fitted within the estimated SRV, and Figures showing results.

Stages First time step event distance from wellbore

SRV growth occurs

Number of events fit in SRV

(%)

Number of events fit in

volume double that of SRV (%)

Figures (4.X)

Down-well direction

(m)

Cross-well direction

(m)

Vertical direction

(m) SRV dimensions interpreted via micro-seismic

7 4 4 1 Yes 27 62 7 - 10 1 20 30 35 No 45 82 11 5 5 50 10 No 27 72 12 8 100 25 8 No 26 65 13 9 20 12 4 Yes 49 75 14 11 8 7 12 Yes 58 90 15

No SRV dimensions interpreted via microseismic 2 60 6 60 No 26 58 16 3 14 70 20 No 9 55 17 4 10 30 10 No 33 48 18 6 0 0 20 Yes 23 61 19 10 15 15 20 Yes 19 52 20 12 10 10 20 Yes 93 100 21

127

Page 145: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.7: Stage 7 SRV length versus width for Case 1.

Fig. 4.8: Stage 7 SRV length versus height for Case 1.

128

Page 146: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.9: Stage 7 SRV width versus height for Case 1.

Fig. 4.10: Stage 7 SRV length versus best fit width (using microseismic data) for Case 1.

129

Page 147: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.11: Stage 1 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

130

Page 148: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.12: Stage 5 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

131

Page 149: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.13: Stage 8 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

132

Page 150: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.14: Stage 9 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

133

Page 151: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.15: Stage 11 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

134

Page 152: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.16: Stage 2 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

135

Page 153: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.17: Stage 3 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

136

Page 154: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.18: Stage 4 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

137

Page 155: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.19: Stage 6 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

138

Page 156: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.20: Stage 10 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

139

Page 157: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.21: Stage 12 SRV dimensions for Case 1: (a) SRV length versus width, (b) SRV length versus best fit width (using microseismic events), (c) SRV length versus height, and (d) SRV width versus height.

The following are some investigation on some stages of first time events that are not reliable

and located away from the horizontal wellbore from Table 4.3:

1. The first time step microseismic events data for Stage 1 are not reliable since there is only

one event that occurred and it is located relatively far away from the horizontal wellbore.

Caffagni and Eaton (2015), in a study of the same well, explained that some events are not

detected because the first time step tensile failure have signal-to-noise ratio (SNR) <<1 and

signal levels <5% of the reference waveform [47].

2. The closest first time step event for Stage 5 and Stage 8 are located outside the cutoff and

towards the East side because it is affected by the depleted zone around an old production

wellbore (microseismic observation wellbore not in operation) on the West side of the

140

Page 158: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

horizontal wellbore. The fracture fluid may increase the pressure within the depleted zone

without creating shear failure and consequent microseismic events.

The differences between the interpreted SRV dimensions from microseismic events with the

estimated SRV dimensions are shown in Figure 4.22a. The average differences for the SRV length,

width, and height are 31%, 24%, and 10%, respectively. On average, 36% of the total microseismic

events are within the estimated SRV as shown in Figure 4.22b.

Fig. 4.22: (a) SRV dimensions differences with microseismic events for six interpreted stages and (b) SRV best fit using microseismic event data for all stages.

The pressure drop along the SRV width is calculated by using the transient pressure drop

equation presented by Ahmed and McKinney (2005) [48]. Shear failure pressure from Ahmed and

McKinney (2005) [48] is compared with the shear failure pressure estimated from finite element

analysis conducted in Chapter 3. The results is shown in Figure 4.23 and the Ahmed and McKinney

(2005) pressure drop equation is given as follows [48]:

Δ𝑃𝑃 = 162.6𝑞𝑞𝑞𝑞𝜇𝜇𝑘𝑘ℎ

�log 𝑡𝑡 + 𝑣𝑣𝑐𝑐𝑔𝑔 � 𝑘𝑘𝜑𝜑𝜇𝜇𝐶𝐶𝑡𝑡𝑟𝑟2

� − 3.23 + 0.87𝑐𝑐� (22)

where ΔP is the pressure drop along the extent of the SRV (psi), q is the injection flow rate (bpd),

B is the formation volume factor for the hydraulic fracture fluid (bbl/stb) (it is assumed to be equal

1), µ is the hydraulic fracture fluid viscosity (cP), k is the reservoir equivalent permeability (mD),

h is the formation thickness (ft.), t is the injection time (hours), φ is the formation porosity fraction),

ct is the total formation compressibility (1/psi), r is the distance (ft.), and s is the formation skin.

141

Page 159: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

(c)

Fig. 4.23: (a) Pressure drop from finite element analysis in Chapter 3, (b) pressure drop derived from transient analysis equation for intact natural fractures, and (c) pressure drop derived from transient analysis equation for open natural fractures.

5

10

15

20

25

30

0 5 10 15 20 25 30

P (M

Pa)

x (m)

Pore Pressure Along SRV Width

t=1s t=330s t=1101s t=2250s

5

10

15

20

25

30

0 5 10 15 20 25 30

P (M

Pa)

x (m)

Pore Pressure Along SRV Width

t=1s t=330s t=1101s t=2250s

142

Page 160: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

A skin of negative seven is assumed based on typically expected hydraulically fractured

reservoir skin values [49]. The equivalent SRV permeability and natural fractures spacing are

calculated from Chapter 3. The transient pressure drop is calculated using Equation (22) only for

Stage 7 with the following assumptions:

1. A low pressure drop is created when the fracture is open. This open fracture is assumed to

be a tensile fracture created by hydraulic fracture or shear fracture along an open natural

fracture. These fractures are assumed to have a high permeability (k=23.4 D) as found in

Chapter 3.

2. A high pressure drop is created by shear fracture created by reopening an intact natural

fracture. This shear fracture is assumed to have a lower permeability (k<23.4 D) that

produces pore pressure equal to the initial reservoir pore pressure (9 MPa). It is assumed

that by the end of reopening the natural fracture, the pore pressure is equal to the initial

reservoir pore pressure.

3. Four time periods are defined to represent different conditions – each time period

represents different failures at different distances and injection times. The first time period

(1 s) represents tensile failure at a distance of 1 m. The second time period (330 s)

represents shear failure at a distance of 6.55 m (first natural fracture). The third time period

(1,101 s) represents shear failure at a distance of 19.65 m (second natural fracture). The

fourth time period (2,250 s) represents shear failure at a distance of 30 m (assumed to be

the boundary of the SRV).

4. Each time period has initial injection BHP. The initial injection BHPs are from Chapter 3.

The first time period has an initial injection BHP of 15.9 MPa. The second time period has

an initial injection BHP of 18.2 MPa. The third time period has an initial injection BHP of

22 MPa. The fourth time period has an initial injection BHP of 27.62 MPa.

5. Equation (22) is used to calculate two cases: (a) intact natural fractures and (b) open natural

fractures.

6. For intact natural fractures case, the natural fractures have lower permeability (less than

23.4 D) and higher pressure drop (shear failure pressure is equal to initial reservoir pressure

of 9 MPa).

143

Page 161: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

7. For open natural fractures case, the natural fractures have higher permeability (equal to

23.4 D) and lower pressure drop (shear failure pressure is higher than initial reservoir

pressure but lower than the initial injection BHP).

8. The permeability along the SRV width for the intact natural fractures case calculated using

Equation (22) are: (a) the first time period assumes k=23.4 D for locations less than 1 m

and k=6.11 D for distance between 1 m and 30 m, (b) the second time period assumes

k=23.4 D for locations less than 6.55 m and k=3.93 D for locations between 6.55 m and 30

m, (c) the third time period assumes k=23.4 D for positions less than 19.65 m and k=3.09

D for distances from 19.65 m to 30 m, and (d) the fourth time period assumes that k=23.4

D for the interval from 0 to 30 m. The permeability is assumed constant from the natural

fracture location to the SRV width boundary. This constant permeability is used to produce

constant pressure from the natural fracture to the SRV width boundary equal to initial

reservoir pressure.

9. The permeability along the SRV width for the open natural fractures case calculated using

Equation (22) are: (a) the first time period assumes k=23.4 D for locations between 0 m

and 1 m and k=6.76 D for distance between 2 m and 30 m, (b) the second time period

assumes k=23.4 D for locations between 0 m and 6.55 m and k=4.22 D for locations

between 10 m and 30 m, (c) the third time period assumes k=23.4 D for positions between

0 m and 19.65 m and k=3.16 D for distances from 23 m to 30 m, and (d) the fourth time

period assumes that k=23.4 D for the interval from 0 to 30 m. The permeability is assumed

constant from the natural fracture location to the SRV width boundary. This constant

permeability is used to produce constant pressure from the natural fracture to the SRV

width boundary equal to initial reservoir pressure.

For intact natural fractures case, the shear failure pressure at 6.55 m and 5.5 minutes (330 s),

equal to initial reservoir pressure, is determined by using Equation (22) as presented in Figure

4.17b. This intact natural fracture case shear failure pressure is lower than the shear failure pressure

of 12.12 MPa obtained from Chapter 3. The smaller shear failure pressure is caused by a higher

pressure drop from the transient pressure drop at intact natural fracture location (lower

permeability) compared to the pressure drop from the analysis in Chapter 3. Moreover, the intact

natural fracture case using the transient pressure drop equation considers a sudden permeability

drop between the open hydraulic fractures and intact natural fractures. The sudden permeability

144

Page 162: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

drop is used to produces the pressure at intact natural fractures equal to the initial reservoir

pressure.

The open natural fracture case, the shear failure pressure at 6.55 m and 5.5 minutes (330 s),

equal to 16.95 MPa, is determined by using Equation (22) as presented in Figure 4.17c. This open

natural fracture case shear failure pressure of 16.95 MPa is higher than the shear failure pressure

of 12.12 MPa obtained from the analysis in Chapter 3. The higher shear failure pressure is caused

by a lower pressure drop from the transient pressure drop at open natural fracture location (high

permeability) compared to the pressure drop from the analysis in Chapter 3. The smaller pressure

drop from open natural fractures case using transient analysis can be caused by Equation (22) does

not consider the pressure drop induced by the rock consolidation.

The intact natural fracture case from Equation (22) results in shear failure at initial reservoir

pressure, which does not occur as shown Mohr-Coulomb failure in Figure 4.6 at the lowest

cohesion (C=0). And the open natural fracture case from Equation (22) produces the shear failure

between initial injection BHP and initial reservoir pressure and it does not consider the pressure

drop induced by the rock consolidation. Therefore, this study uses the shear failure pressure

obtained from Chapter 3 to generate the Mohr circle on the Mohr-Coulomb failure envelope

(Chapter 3 considers pressure drop induced by the rock consolidation).

The estimated SRV dimensions of some stages have differences with the interpreted SRV

dimensions via microseismic. For example, the estimated SRV length of Stage 9 is 36% longer

than the microseismic interpretation. From the microseismic observation, the estimated SRV

length fits better with the results estimated from the microseismic events. The estimated SRV

covers 49% of total microseismic events. The estimated SRV widths of Stages 8 and 9 are 56%

and 47% shorter than the interpreted SRV width estimated from the microseismic data. This

phenomenon might be caused by the effect of an old depleted zone around the Stages 8 and 9.

There is an old wellbore on the North East side of the Stages 8 and 9 which has been producing

since 1982. Hydraulic fracturing of new wells in the vicinity of the old well might create shear

failures along natural fractures triggering additional non-productive microseismic events. Maxwell

et al. (2002) discovered that the hydraulic fractures grew toward old neighboring wells because of

the depleted zones around the older wells [21]. Maxwell et al. (2010) showed that these cases

occurred when critically stressed fractures existed in the neighborhood of the hydraulic fracture

145

Page 163: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

failures [21]. These critically stressed fractures could trigger small stresses changes and result in

remote triggering of microseismic events [21]. Therefore the wider SRV by these trigger events

might not have any connection to the SRV dimensions of Stages 8 and 9. Furthermore, these

reopening natural fractures might not contribute to the horizontal wellbore production in the future.

The SRV length of Stage 5 is 47% shorter than the interpreted microseismic events. The

microseismic observations shows that there are two clouds of second time step events around Stage

5. The dimensions of the SRV are estimated by using the first cloud of second time step events.

The closest second cloud event from the first time step event center is located at 270 m. Therefore

it is assumed that the first time step and the second time step cloud are not connected and that the

estimated SRV length of Stage 5 fits better with the microseismic events.

The average estimated SRV length, width and height are 166.30 m, 56.83 m, and 62.0 m for

all the 12 stages as listed in Table 4.4. The SRV dimensions are a complex fracture network (they

are not bi-wing fracture) because of the low ratio between the maximum horizontal stress with the

overburden stress. Weng et al. (2011), from their simulations, showed that the stress ratio, natural

fractures, and internal friction angle affected the complexity of the fracture network [25].

Specifically, their results showed that lowering the stress ratio changed the system from a bi-wing

fracture to a complex fracture network.

This study suggests an optimum fracture spacing that will drain all the area (SRV) between the

injection ports. An optimum fracture spacing is assumed to be uniform and equal to the average

estimated SRV width (56.83 m), the interpreted SRV dimensions from microseismic were only

done for six stages. It is shorter than the hydraulic fracture injection port spacing (116.82 m) listed

in Table 4.5. Therefore, it is suggested that the hydraulic fracture injection port spacing should be

shorter and equal to that of the SRV width to improve drainage from the reservoir.

146

Page 164: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 4.4: The estimated SRV dimensions results and the located microseismic events.

Stages SRV length (m) SRV width (m) SRV height (m) Number of events

1 192.93 66.70 80.42 38 2 173.02 65.10 61.99 35 3 165.96 55.61 58.43 33 4 162.33 53.27 60.49 92 5 164.66 56.46 61.12 109 6 162.59 51.98 55.96 59 7 174.00 60.00 60.00 46 8 160.30 46.96 54.68 124 9 159.53 49.59 54.59 72

10 155.54 48.70 52.22 27 11 149.65 56.81 60.21 81 12 175.10 70.81 84.17 16

Average 166.30 56.83 62.02

Table 4.5: Comparison of SRV width and hydraulic fracture port spacing.

Stages SRV width (m) Stages Hydraulic Fracture Injection Port spacing (m)

1 66.70 1-2 81.30 2 65.10 2-3 108.84 3 55.61 3-4 109.49 4 53.27 4-5 120.84 5 56.46 5-6 121.63 6 51.98 6-7 133.62 7 60.00 7-8 122.09 8 46.96 8-9 121.52 9 49.59 9-10 109.92 10 48.70 10-11 145.92 11 56.81 11-12 109.90 12 70.81 Average port spacing

(m) 116.82

Average 56.83

Xu et al. (2009) and Weng et al. (2011) had suggested that the ratio of horizontal stresses

affected the SRV length and width [22, 25]. Here, the ratio of horizontal stresses is equal to 2.21

and the average ratio of SRV length to height is 2.95. The differences between these two ratios

might depend on the hydraulic diffusivity coefficients in the maximum and minimum horizontal

147

Page 165: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

stress directions. The average ratio of diffusivity coefficients in the maximum and minimum

horizontal stress directions is equal to 1.34. Darcy’s law and the mass continuity equation show

that fracture fluid propagation depends on hydraulic diffusivity and the proposed model for SRV

dimensions show that the SRV dimension depends on the stress. Therefore, an equation is

proposed based on Darcy’s law, mass continuity equation, and proposed SRV model:

𝑡𝑡𝑏𝑏

= 𝜎𝜎𝐻𝐻𝜎𝜎ℎ

𝜂𝜂�𝐻𝐻𝜂𝜂�ℎ

(23)

where a and b is the SRV length and width, �̅�𝜂𝐻𝐻 and �̅�𝜂ℎ are the equivalent hydraulic diffusivity

coefficient in maximum and minimum horizontal stresses directions. This implies:

𝜂𝜂�𝐻𝐻𝜂𝜂�ℎ

=𝑘𝑘�𝐻𝐻𝐶𝐶�𝐻𝐻𝑘𝑘�ℎ𝐶𝐶�ℎ

(24)

where 𝑘𝑘�𝐻𝐻 and 𝑘𝑘�ℎ are the equivalent permeability in maximum and minimum horizontal stresses

directions and 𝑐𝑐�̅�𝐻 and 𝑐𝑐ℎ̅ are the equivalent compressibility in maximum and minimum horizontal

stresses directions defined as [50]:

𝑐𝑐 = 1𝐾𝐾

(25)

𝐾𝐾 = 𝜋𝜋3(1−2𝑣𝑣) (26)

𝑐𝑐 = 3(1−2𝑣𝑣)𝜋𝜋

(27)

where 𝐸𝐸 is the equivalent Young’s modulus, also 𝑣𝑣 is the equivalent Poisson’s ratio, K is the bulk

modulus. Similarly, Yu and Aguilera (2012) also proposed that the SRV dimensions depended on

the hydraulic diffusivity [26]. The results of the diffusivity ratios in maximum and minimum

horizontal stresses directions are listed in Table 4.6.

148

Page 166: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 4.6: Diffusivity coefficient ratios for maximum and minimum horizontal stresses direction.

Stages SH (psi) Sh (psi) SRV Length/SRV Width SH/Sh ηH/ηh 1 7104.48 3212.66 2.89 2.21 1.31 2 7104.48 3212.66 2.66 2.21 1.20 3 7104.48 3212.66 2.98 2.21 1.35 4 7104.48 3212.66 3.05 2.21 1.38 5 7104.48 3212.66 2.92 2.21 1.32 6 7104.48 3212.66 3.13 2.21 1.41 7 7104.48 3212.66 2.90 2.21 1.31 8 7104.48 3212.66 3.41 2.21 1.54 9 7104.48 3212.66 3.22 2.21 1.45 10 7104.48 3212.66 3.19 2.21 1.44 11 7104.48 3212.66 2.63 2.21 1.19 12 7104.48 3212.66 2.47 2.21 1.12

Average 2.95 2.21 1.34 4.5.2 Total Stress Change

The first results in the previous section are produced assuming constant total stresses (Case 1).

The results in this section are produced using total stress changes (Case 2). The total stress change

and the pore pressure for Case 2 are estimated from Chapter 3. The changes in the total stress affect

the Mohr circle in the Mohr-Coulomb failure envelope (Figure 4.24). From Chapter 3, the total

stresses (both the maximum and the minimum horizontal stress) increase with the injection time.

In Case 2, it is assumed only the minimum horizontal stress change with the maximum horizontal

stress and the overburden stress are constant. Therefore only the width of the SRV changes (SRV

length and height constant).

149

Page 167: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 4.24: Mohr-Coulomb stress failure envelope of stage 7 for Case 2.

The first time step of Case 2 uses the minimum horizontal stress increases of 4.54 MPa

(minimum horizontal stress value at 1 s and located at the wellbore from Chapter 3), BHP of 41.5

MPa (fracture breakdown pressure) and fracture breakdown pressure duration of 5.5 minutes. The

second time step of Case 2 uses the minimum horizontal stress increase of 0.9 MPa (minimum

horizontal stress value at 330 s and located at 6.55 m from Chapter 3), and pressure of 12.12 MPa,

and fracture propagation duration (from 5.5 minutes after hydraulic fracture starts which is the

start of hydraulic fracture propagation until the end of injection).

Case 1 and Case 2 have naturally fractured rock cohesion equal to 0.69 MPa. Case 2 SRV

width is shorter by 4.11% compared to the Case 1 SRV width. This suggests that the SRV

dimensions are overestimated when total stresses changes are not considered. Mcclure and Horne

(2013) studied the effect of stress changes during hydraulic fracturing and explained the

importance of including the stresses induced by the deformation in hydraulic fracture modeling

[28]. These stresses directly impact hydraulic fracture propagation and resulting fracture network

properties [28]. The Mohr-Coulomb failure envelope properties and plot for the Case 2 are

presented in Table 4.7.

150

Page 168: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 4.7: Stage 7 Mohr-Coulomb failure envelope properties for Case 2.

Mohr-Coulomb failure envelope properties

Tensile failure (not shown in Mohr-Coulomb

failure envelope

Shear failure (red Mohr circle)

Effective Sh (psi) -2140.31 1593 Effective Sh (MPa) -14.76 10.88 Effective SH (psi) 1088 5349 Effective SH (MPa) 7.5 36.88 Additional Sh (MPa) 4.54 0.9 Additional Sh (psi) 658 131 Additional SH (MPa) 0 0 Additional SH (psi) 0 0 Pore pressure cause failure (MPa)

41.5 12.12

Pore pressure cause failure (psi) 6019 1758

Table 4.8 compares the Mohr-Coulomb failure envelope properties for Case 1 and Case 2. To

support the shear failure determined by using the Mohr-Coulomb failure envelope, the effective

stresses are plotted for the intact rock and the naturally fractured reservoir based on the study done

by Hoek and Martin (2014) (Figure 4.25) [51]. For intact rock, tensile strength is calculated by

using the Stage 7 fracture breakdown BHP and tensile strength is assumed to be zero for naturally

fractured reservoir. The UCS is calculated in Section 4.4 with no shear failure data. Therefore for

intact rock, the shear failure envelope is assumed to be a straight line extension from the tensile

strength and UCS.

151

Page 169: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 4.8: Stage 7 SRV dimensions and Mohr-Coulomb failure envelope properties comparison Cases 1 and 2.

Stage 7 Case 1 Case 2 Differences %

SRV width (m) 60 58 4 Increase Sh at tensile failure (MPa) 4.5 Increase Sh at tensile failure (psi) 658.9 Tensile failure Sh (MPa) -19.3 -14.8 Tensile failure Sh (psi) -2799.2 -2140.3 Increase Sh at shear failure (MPa) 0.9 Increase Sh at shear failure (psi) 130.9 Shear failure Sh (MPa) 10.1 11 Shear failure Sh (psi) 1462 1593 BHP shear failure (MPa) 12.12 12.12 BHP shear failure (psi) 1758 1758 Cohesion (MPa) 0.69 0.69 Cohesion (psi) 100 100

Fig. 4.25: Plot of Case 1 and Case 2 effective maximum horizontal stress versus effective minimum horizontal stress for intact rock and naturally fractured reservoirs.

152

Page 170: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The UCS of 2.38 MPa is calculated by using the estimated cohesion for the naturally fractured

reservoir. The naturally fractured reservoir for Case 1 and Case 2 only has one failure point that

touches the failure envelope. The failure point for the cases have pore pressure of 12.12 MPa. The

failure point for Case 1 and Case 2 have the same effective maximum horizontal stress and

different effective minimum horizontal stress. Case 2 has a higher effective minimum horizontal

stress of 11 MPa compared to the Case 1 effective minimum horizontal stress of 10.1 MPa. The

failure line between the UCS and failure point for the Cases 1 and 2 have small differences in

inclination (0.45o) which is negligible. This plot supports that the initial reservoir pressure does

not experience the shear failure yet (not touching the failure line for Case 1 or Case 2). The SRV

dimensions results for the minimum horizontal stress case are shown in Figure 4.26.

Fig. 4.26: SRV dimensions of Stage 7 for Case 2 SRV dimensions: (a) SRV length versus width, (b) SRV length versus width best fit with microseismic events, (c) SRV length versus height, and (d) SRV width versus height.

153

Page 171: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.6 Conclusions

New models to estimate the dimensions of SRV are derived and they can be used to optimize

the design of future hydraulic fracture jobs. It is found that the SRV dimensions can be related to

the effective stresses, injected fluid volume, and plane strain modulus as well as the naturally

fractured reservoir cohesion strength and internal friction angle. The models are calibrated with

SRV dimensions interpreted via microseismic events from 6 hydraulic fracture stages. The models

have the ability to estimate the evolution of SRV dimensions during hydraulic fracturing by using

microseismic events center of two consecutive time periods. The SRV dimensions represent the

extent of the fracture network created by the multi-stage hydraulic fracture in a horizontal well in

a naturally fractured reservoir.

The conclusions of this chapter are as follows:

1. Two different time periods with each pressure and injection time are required to represent

tensile and shear failure. The first time step that experiences tensile failure has BHP that is

measured from field fracture breakdown pressure and it is assumed there is no pressure

drop from the horizontal wellbore to the tensile failure location. The second time step that

experiences shear failure has no pore pressure measured at the field. It is assumed that shear

failure occurs at natural fracture location and start of fracture propagation until the end of

injection.

2. Growth of the SRV is not apparent for some of the hydraulic fracture stages. In some stages,

the first time step events are not triggered in the proximity of the horizontal wellbore. This

phenomenon could be caused by the first time step events were not detected by the receiver.

And also, it could be caused due to low signal to noise ratio and signal levels compare to

reference waveform signal. Another reason to this phenomenon as because the events were

affected by depleted zone at old production wellbore on the West side of the horizontal

wellbore. The fracture fluid may be increasing the pressure within the depleted zone.

3. The impact of the total stress changes on the dimensions of the SRV reveals that total

stresses must be included to avoid overestimation of the dimensions of the SRV. The total

stress changes affect the Mohr circle in the Mohr-Coulomb failure envelope.

154

Page 172: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

4.7 References

[1] BP Statistical Review. 2015. http://www.bp.com/content/dam/bp/pdf/Energy-economics/statistical-review-2015/bp-statistical-review-of-world-energy-2015-full-report.pdf (Downloaded 1 September 2015).

[2] Fjaer, E., Holt, R. M., Horshud, P., Raaen, A. M., and Risnes, R. 2008. Petroleum Related Rock Mechanics, second edition. Amsterdam: Elsevier B. V.

[3] Lockner, D. A. 1995. Rock Failure. In Rock Physics and Phase Relations, first edition, T.J. Ahrens, Chap. 8, 127-147. Washington, D. C.: American Geophysical Union.

[4] Zoback, M. D. 2007. Reservoir Geomechanics, first edition. Cambridge: Cambridge University Press.

[5] Jaeger, J. C., and Cook, N. G. W. 1979. Fundamentals of Rock Mechanics, third Edition. London: Chapman and Hall.

[6] Warpinski, N. R., and Branagan, P. T. 1988. Altered Stress Fracture. Presented at the 1988 Rocky Mountain Regional Meeting, Casper, USA, 11-13 May. SPE-17533. http://dx.doi.org/10.2118/17533-PA

[7] Jupe, A. J., Willis-Richards, J., and Nicollis, J. D. M. 1993. Review of HDR projects. CSM Associates Ltd., Report 1R03/06 for AEA Technology Ltd. Harwell, U.K.

[8] Duchane, D. 1998. The History of HDR Research and Development. Proc., the Fourth International HDR Forum, Strasbourg, France, 28-30 September, LA-UR--98-4067; CONF-9809136. http://www.osti.gov/geothermal/biblio/350838

[9] Rahman, M. K., Hossain, M. M., and Rahman, S. S. 2002. A Shear-Dilation Based Model for Evaluation of Hydraulically Stimulated Naturally Fractured Reservoirs. Int. J. Num. Analyt. Methods Geomech. 26:469-497.

http://dx.doi.org/10.1002/nag.208 [10] Palmer, I., Moschovidis, Z., and Cameron, J. 2007. Modeling Shear Failure and Stimulation

of the Barnett Shale After Hydraulic Fracture. Presented at the 2007 Hydraulic Fracture Technology Conference, Texas, USA, 29-31 January. SPE-106113. http://dx.doi.org/10.2118/106113-MS

[11] Khristianovitch, S. A. and Zheltov, Y. P. 1955. Formation of Vertical Fractures by Means of Highly Viscous Fluids. Presented at the 4th World Petroleum Congress, Rome, Italy, 6-15 June. WPC-6132.

[12] Geertsma, J., and de Klerk, E. 1969. A Rapid Method of Predicting Width and Extent of Hydraulically Induced Fractures. J. Pet. Tech. 21: 1571-1581.

[13] Perkins, T. K. and Kern, L. R. 1961. Widths of Hydraulic Fractures. J Pet Tech 13 (9): 937-949. SPE-89-PA.

http://dx.doi.org/10.2118/89-PA [14] Rahman, M. M. and Rahman, M. K. 2010. A Review of Hydraulic Fracture Models and

Development of an Improved Pseudo-3D Model for Stimulating Tight Oil/Gas Sand. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects 32 (15): 1416-1436.

http://dx.doi.org/10.1080/15567030903060523

155

Page 173: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[15] Barree, R. D. 2009. Modeling Fracture Geometry. St. Petersburg State Polytechnical University, Institute of Applied Mathematics and Mechanics. http://mech.spbstu.ru/images/4/4d/Modeling_Fracture_Geometry.pdf (Downloaded 24 March 2015).

[16] Sneddon, I. N. 1946. The Distribution of Stress in the Neighborhood of a Crack in an Elastic Solid. Proc. The Royal Society of London. Series A: Mathematical and Physical Sciences 187 (1009): 229-260.

http://dx.doi.org/10.1098/rspa.1946.0077 [17] Nordgren, R. P. 1970. Propagation of a Vertical Hydraulic Fracture. SPE J 12 (04): 306-

314. http://dx.doi.org/10.2118/3009-PA

[18] Daneshy, A. A. Jr. 1971. Opening of a Pressurized Fracture in an Elastic Medium. Presented at the 22nd Annual Technical Meeting of the Petroleum Society of CIM, Banff, Canada, 2-4 June. PETSOC-7616.

http://dx.doi.org/10.2118/7616 [19] Economides, M. J. and Nolte, K. G. 2000. Reservoir Stimulation, third edition. West

Sussex: John Wiley & Sons Ltd. [20] Fisher, M. K., Wright, C. A., Davidson, B. M., Goodwin, A. K., Fielder, E. O., Buckler, W.

S., and Steinsberger, N. P. 2002. Integrating Fracture Mapping Technologies to Optimize Stimulations in the Barnett Shale. Presented at the SPE Annual Conference and Exhibition, San Antonio, Texas, 29 September–2 October. SPE-77441.

http://dx.doi.org/10.2118/77441-MS [21] Maxwell, S. C., Urbancic, T. I., Steinsberfer, N., and Zinno, R. 2002. Microseismic Imaging

of Hydraulic Fracture Complexity in the Barnett Shale. Presented at the SPE Annual Conference and Exhibition, San Antonio, Texas, 29 September–2 October. SPE-77440. http://dx.doi.org/10.2118/77440-MS

[22] Xu, W., Calvez, J. L., and Thiercelin, M. 2009. Characterization of Hydraulically-Induced Fracture Network Using Treatment and Microseismic Data in a Tight-Gas Formation: A Geomechnical Approach. Presented at the 2009 SPE Tight Gas Completions Conference, San Antonio, 15-17 June. SPE-125237.

http://dx.doi.org/10.2118/125237-MS [23] Maxwell, S. C., Underhill, B., Bennet, L., and Martinez, A. 2010. Key Criteria for a

Successful Microseismic Project. Presented at the SPE Annual Technical Conference and Exhibition. Florence, Italy, 19–22 September. SPE-134695. http://dx.doi.org/10.2118/134695-MS

[24] Mayerhofer, M. J., Lolon, E. P., Warpinski, N. R., Cipolla, C. L., Walser, D., and Rightmire C. M. 2010. What is stimulated reservoir volume? SPE Prod & Oper 25(01): 89-98. SPE-119890. http://dx.doi.org /10.2118/119890-PA

[25] Weng, X., Kresse, O., Cohen, C., Wu, R., and Gu, H. 2011. Modeling of Hydraulic-Fracture-Network Propagation in a Naturally Fractured Formation. SPE Prod & Oper 26(04): 368-380. SPE-140253.

http://dx.doi.org/10.2118/140253-PA

156

Page 174: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[26] Yu, G., and Aguilera, R. 2012. 3D Analytical Modeling of Hydraulic Fracture Stimulated Reservoir Volume. Presented at the SPE Latin American and Caribbean Petroleum Engineering Conference, Mexico City, Mexico, 16-18 April. SPE-153486. http://dx.doi.org/10.2118/153486-MS

[27] Nassir, M., Settari, A. and Wan, R. 2012. Prediction of SRV and Optimization of Fracture in Tight Gas and Shale Using a Fully Elasto-Plastic Coupled Geomechanical Model. Presented at the SPE Hydraulic Fracture Technology Conference, The Woodlands, Texas, 4-6 February. SPE-163814

http://dx.doi.org/10.2118/163814-PA [28] McClure, M. W. and Horne, R. N. 2013. Discrete Fracture Network Modeling of Hydraulic

Stimulation Coupling Flow and Geomechanics, first edition. Heidelberg, Germany: Springer International Publishing.

[29] ESG Solutions, 2015, Microseismic Event Location. https://www.esgsolutions.com/technical-resources/microseismic-knowledgebase/microseismic-event-location (Downloaded 20 May 2015).

[30] Raleigh, C. B., Healy, J. H., and Bredehoeff, J. D. 1976. An Experiment in Earthquake Control at Rangely, Colorado. Science 191 (4233): 1230-1237.

http://dx.doi.org/10.1126/science.191.4233.1230 [31] Vinegar, H. J., Wills, P. B., De Martini, D. C., Shlyapobersky, J., Deeg, W. F. J., Adair, R.

G., Woerpel, J. C., Fix, J. E., and Sorrells, G. G. 1992. Active and Passive Seismic Imaging of a Hydraulic Fracture in Diatomite. J Pet Technol 44 (01): 28-90. SPE-22756-PA. http://dx.doi.org/10.2118/22756-PA

[32] Withers, R. J., Perkins, T. K., and Keck, R. G. 1996. A Field Demonstration of Hydraulic Fracture for Solids Waste Disposal, Part Two: Real Time Passive Seismic Monitoring System and Data Analysis. In Deep Injection Disposal of Hazardous and Industrial Waste, J. A. Apps and C. F. Tsang, San Diego: Academic Press.

[33] Warpinski, N. R., Branagan, P. T., Peterson, R. E., Wolhart, S. L., and Uhl, J. E. 1998. Mapping Hydraulic Fracture Growth and Geometry Using Microseismic Events Detected by a Wireline Retrievable Accelerometer Array. Presented at the 1998 Gas Technology Symposium, Calgary, 15-18 March. SPE-40014. http://dx.doi.org/10.2118/40014-MS

[34] Griffin, L. G., Sullivan, R. B., Wolhart, S. L., Waltman, C. K., Weijers, L., and Warpinski, N. R. 2003. Hydraulic Fracture Mapping of the High-Temperature, High-Pressure Bossier Sands in East Texas. Presented at the 2003 SPE Annual Technical Conference and Exhibition, Denver, 5-8 October. SPE-84489. http://dx.doi.org/10.2118/84489-MS

[35] Wright, C. A. 1998. Tiltmeter Fracture Mapping From the Surface and Now Downhole. Petr Eng Int 71 (1): 50-54. http://ezproxy.lib.ucalgary.ca/login?url=http://search.proquest.com/docview/224546974?accountid=9838

[36] Warpinski, N. R., Griffin, L. G., and Davis, E. J. 2006. Improving Hydraulic Fracture Diagnostics by Joint Inversion of Downhole Microseimic and Tiltmeter Data. Presented at the 2006 SPE Annual Technical Conference and Exhibition, San Antonio, 24-27 September. SPE-102690. http://dx.doi.org/10.2118/102690-MS

157

Page 175: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[37] Hall, M., and Kilpatrick, J. E. 2009. Surface Microseismic Monitoring of Slick-Water and Nitrogen Fracture Simulations, Arkoma Basin, Oklahoma. Presented at the SEG Houston 2009 International Exposition and Annual Meeting. SPE-132371-MS. http://dx.doi.org/10.2118/132371-MS

[38] Barker, W. 2009. Increased Production Through Microseismic Monitoring of Hydraulic Fracture Over a Multiwell Program. Presented at the 2009 SPE Annual Conference and Exhbition, New Orleans, Lousiana, 4-7 October. Paper SPE 124877.

http://dx.doi.org/10.2118/124877-MS [39] Davis, C. 2012. Report on ConocoPhillips Hoadley Microseismic Monitoring on Well A.

Technical report. ESG Solutions. Canada (unpublished). [40] Bell, J. S., and S. Bachu. 2003. In situ Stress Magnitude and Orientation Estimates for

Cretaceous Coal-Bearing Strata Beneath the Plains Area of Central and Southern Alberta. Bulletin of Canadian Petroleum Geology 51(1): 1-28.

http://dx.doi.org/10.2113/gscpgbull.51.1.1 [41] Hawkes, C. 2013. An Overview of Rock Mechanical Factors Affecting Petroleum

Development in The Williston Basin. Presented at the Williston Basin Petroleum Conference, Regina, Canada, May.

[42] Rich, J. P. and M. Ammerman, M. 2010. Unconventional Geophysics for Unconventional Gas. Presented at the SPE Unconventional Gas Conference, Pittsburgh, USA, 23-25 February.

http://dx.doi.org/10.2118/131779-MS [43] Bratton, T. 2011. Hydraulic Fracture Complexity and Containment in Unconventional

Reservoirs. Oral presentation given at the ARMA Workshop on Rock Mechanics/Geomechanics, San Fransisco, USA, 23 June.

[44] Hareland, G. and Nygaard, R. 2007. Calculating Unconfined Rock Strength from Drilling Data. Presented at the 1st Canada – U.S. Rock Mechanics Symposium, Vancouver, Canada, 27-31 May. ARMA-07-214.

[45] Goodarzi, S. and Settari, A. 2009. Geomechanical Modelling and Analysis Wabamun Area CO2 Sequestration Project (WASP). Prepared for Energy and Environmental Systems Group, Calgary, Canada, September.

[46] Hoek, E. 1983. Strength of Jointed Rock Masses. Geotechnique 23 (3): 187-223. [47] Caffagni, E. and Eaton, D. 2015. Detection and Location of Small Microseismic Events

Using a Matched Filtering Algorithm (MFA). Presented at the Geoconvention 2015, Calgary, Alberta, Canada, 4-8 May.

[48] Ahmed, T, and McKinney, P. D. 2005. Advanced Reservoir Engineering, first edition. Massachusetts: Elsevier Inc.

[49] Fekete Webhelp. 2015. http://www.fekete.com/san/webhelp/welltest/webhelp/Content/HTML_Files/Reference_Materials/Skin.htm (Downloaded 20 May 2015).

[50] Zimmerman, R. W., Somerton, W. H., and King, M. S. 1986. Compressibility of Porous Rock. Journal of Geophysical Research 91 (B12): 12,765-12,777.

[51] Hoek, E., and Martin, C. D. 2014. Fracture Initiation and Propagation in Intact Rock – A Review. Journal of Rock Mechanics and Geotechnical Engineering 6 (4): 287-300. http://dx.doi:10.1016/j.jrmge.2014.06.001

158

Page 176: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

CHAPTER 5: DETERMINATION OF FRACTURE

CHARACTERISTICS WITHIN STIMULATED ROCK

VOLUME USING DIFFUSIVITY EQUATION AND

PRODUCTION ANALYSIS

Summary

Multifractured horizontal wells have been widely used to produce unconventional reservoirs since

the 1980s. This type of completion and stimulation creates a complex fracture network in naturally

fractured reservoirs and enhances the drainage area and permeability of the near-well region. The

complex fracture network, also called stimulated rock volume (SRV), enables commercial

production of gas from unconventional reservoirs. Determination of SRV properties is important

to confirm the effectiveness of the hydraulic fracturing job. In this study, production analysis is

used to determine the SRV properties from a tight gas reservoir of Glauconitic Formation in the

Hoadley Field, Alberta, Canada. A linear flow regime is identified during an 8 months production

interval using finite conductivity fracture type curves. The history match is done by using the

horizontal well multifractured enhanced fracture region with dual porosity model in a rate transient

analysis simulator. The RTA results are compared with the results from the diffusivity equations.

The SRV permeability during production from the RTA and the developed diffusivity equations

are compared with the permeability during the injection from finite element analysis. This study

gives an insight regarding the SRV permeability changes from injection to production.

5.1 Introduction

Tight gas reservoirs are a great source of hydrocarbon in an era of declining conventional

reservoir production. This type of reservoir has very low permeability and porosity which requires

stimulation and completion such as that realized by using multifractured horizontal wells to be

produced efficiently. Most tight gas reservoirs are naturally fractured. Hydraulic fracturing a

naturally fractured reservoir leads to a complex fracture network, also referred to as the stimulated

159

Page 177: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

rock volume (SRV). The efficiency of the hydraulic fracturing job and design can be predicted by

determining the SRV permeability during injection (creation of the hydraulic fractures) and

production. The SRV permeability is also crucial in the long term production forecast. Despite the

importance of the precision of SRV permeability’s prediction, there are not many studies that have

been done on this topic and specifically in naturally fractured reservoir in the Glauconitic

Formation in the Hoadley Field, Alberta, Canada. In this study, the SRV permeability will be

determined by using production data fitted to a newly developed diffusivity equation.

5.1.1 Objective of Study

The objective of this study is to estimate the characteristics of the stimulated rock volume

during production by using both diffusivity equation and rate transient analysis, specifically the

permeability, pressure and pressure drop gradient profile, and porosity within the SRV. The gas

flow rate is calculated from Darcy’s law which in turn requires the SRV permeability, pressure

gradient, gas viscosity, and wellbore area. The fracture compressibility (assumed constant) is also

needed to match the initial gas flow rate. It is assumed that the total compressibility is given by

the fracture compressibility (fluid and formation compressibility are small and neglected). The

calculated gas flow rate profile is then compared with the field gas flow rate profile. The simulated

SRV permeability from the diffusivity equation and rate transient analysis are compared.

5.2 Literature Review

5.2.1 Behavior of Naturally Fractured Reservoir

Bulnes and Fitting (1945) and Imbt and Ellison (1946) differentiated the types of porosities in

the rocks [1, 2]. The primary porosities are intergranular where they are controlled by deposition

and lithification and they are highly interconnected. Void systems of sandstones are typical of the

primary porosity. The secondary porosity is small in openings and it is controlled by fracturing or

jointing where it is not highly interconnected. These types of porosity can be channels or vugular

voids that had been developed during weathering or burial such as limestones or dolomites. Joints

or fissures are another types of secondary porosities in shale, siltstone, limestone or dolomite and

160

Page 178: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

they are usually vertical where shrinkage cracks are resulting from chemical process and don’t

have any preferred direction. In most cases, the two types of porosities are found together in the

rock. The realization of this porous medium is as a complex of discrete volumetric elements with

anisotropic primary porosities coupled with secondary porosities as shown in Figure 2.14 in

Chapter 2.

Warren and Root (1963) were the first to propose that a reservoir could contain both primary

(intergranular) and secondary (fissure or vugular) porosities [3]. They assumed the primary

porosity region contributes significantly to the pore volume but contributes insignificantly to the

flow capacity. They developed an idealized model to study the behavior of dual porosity system

(naturally fractured reservoir) with pseudo steady fluid transfer from matrix to fracture. Their study

proposed two parameters to describe the deviation of the behavior of dual porosity medium from

a homogeneous porous medium. The first parameter is ω. It is a measure of the fluid capacity of

the secondary porosity or it is also defined as storativity ratio or a fraction of the total pore volume

associated with one of the porosities. The second parameter is λ. It is the heterogeneity scale that

is present in the medium or it is also called as interporosity flow coefficient or the ratio of the

permeability of the matrix to the permeability of the fractures. The two parameters are shown

below:

𝜔𝜔 = 𝜙𝜙𝑓𝑓𝐶𝐶𝑓𝑓𝜙𝜙𝑓𝑓𝐶𝐶𝑓𝑓+𝜙𝜙𝑚𝑚𝐶𝐶𝑚𝑚

(1)

𝜆𝜆 = 𝛼𝛼𝑟𝑟𝑤𝑤2𝑘𝑘𝑚𝑚𝑘𝑘𝑓𝑓

(2)

where 𝑡𝑡𝑓𝑓 and 𝑡𝑡𝑚𝑚are fracture and matrix porosity, 𝑐𝑐𝑓𝑓 and 𝑐𝑐𝑚𝑚 are fracture and matrix

compressibility, α is shape factor, rw is wellbore radius, kf and km are fracture and matrix

permeability.

Stearns (1982) defined the natural fracture as a macroscopic planar discontinuity that resulted

from stresses that exceeded the rupture strength of the rock [4]. Nelson (1985) defined another

definition of natural fracture as a naturally occurring macroscopic planar discontinuity in rock due

to deformation or physical diagenesis and the natural fractures effects could be positive or negative

on the fluid flow [5]. Aguilera (1998) explained that virtually all reservoirs contained at least some

161

Page 179: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

natural fractures but if the effect of the natural fractures could be neglected then the reservoir could

be classified into conventional reservoir [6].

Therefore, the reservoir that contains natural fractures is called naturally fractured reservoir.

Most of the reservoirs have natural fractures, yet the effects of natural fractures are not quite

understood and mostly underestimated. The natural fractures are the main production factor in a

wide range of unconventional reservoir including tight gas reservoir.

Aguilera (2003) also mentioned that it was important to know the magnitude and the

orientation of the in-situ stresses, spacing, aperture, permeability and porosity of the fractures, also

permeability and porosity of the matrix [7]. Stearns (1982), Nelson (1985), and Aguilera (1998)

classified the natural fractures from the geological point of view as tectonic (fold or fault related),

regional, contractional (diagenetic) and surface related [4, 5, 6].

McNaughton and Garb (1975), Aguilera (1995), and Aguilera (2003) classified the naturally

fractured reservoirs from a storage point of view as Type A, B or C [7, 8, 9]. Type A had a large

amount of hydrocarbon stored in the low permeability matrix and small amount stored in the high

permeability fractures. Type B had half of the hydrocarbon stored in the matrix and half stored in

the fractures. Type C had all the hydrocarbon stored in the fractures. Aguilera (1998) showed

some range of recovery for different types of naturally fractured gas reservoir and recovery

mechanism [6].

Aguilera (2003) explained about the engineering aspects of the natural fractures identification

quantitatively based on geophysics, geology and engineering [7]. Aguilera (1995) mentioned that

the identification could be done by using the direct and the indirect methods [9]. The direct

methods include the core measurement, the drill cuttings, the downhole photographs and the

borehole videos. The indirect methods are the outcrops, the drilling history (loss circulation

information during drilling operation), the well logs and the seismic measurement [9]. Another

important properties of the naturally fractured reservoir is the fracture compressibility. Aguilera

(2003) stated that the fracture compressibility for zero mineralization within the fracture, should

be higher than the matrix compressibility because of the unrestricted fluid flow [6]. The differences

between these compressibility values depend on the amount of the secondary mineralization within

the fractures, the fracture orientation, and the in-situ stresses also the reservoir pressure condition

[6]. Since there is no core available to determine the fracture compressibility, the correlation

162

Page 180: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

proposed by Aguilera (1998) will be used in this research to determine the fracture compressibility

(it is illustrated in Figure 2.15 in Chapter 2) [7].

5.2.2 Dual Porosity Model (Pseudo Steady State and Transient)

Brown et al (2009) explained that there were two common models of dual porosity idealization

[10]. There was a model that considered pseudo steady fluid transfer from matrix to fracture as the

pseudo steady model described by Warren and Root (1963) [3]. The other model considered

transient fluid transfer from matrix to fracture referred to as the transient model described

byKazemi (1969), de Swan-O (1976) and Serra et al., (1983) [11, 12, 13]. The transient models

were divided into slabs, cubes and sticks model.

Wattenbarger et al. (1998) developed a model for a vertically fractured well at the center of a

rectangular dual porosity reservoir by considering a slightly compressible fluid with constant

viscosity in a rectangular reservoir with closed outer boundaries [14]. Initial pressure was uniform

throughout the reservoir and the model did not consider the skin and wellbore storage.

Wattenbarger et al. (1998) introduced a mathematical model describing the linear transient dual

porosity reservoir.

5.2.3 Flow Regimes of Multi-Fractured Horizontal Well in Naturally

Fractured Reservoir

Chen and Raghavan (1997) proposed the flow regimes for a multifractured horizontal well in

a rectangular drainage region for two fractures [15]. They neglected the wellbore storage effect.

The first flow regime was bilinear or linear flow [15]. The bilinear flows occurred when the

fracture conductivity was finite and the fracture length was greater than the fracture height [15].

The linear fluid flow within the fracture towards the horizontal well and within the formation is

shown in Figure 2.16 in Chapter 2 ([16].

Nobakht et al. (2011) stated that the second flow regime was early linear flow that occurred

from the formation toward the fractures and the flow within the fractures was negligible [17]. In

multifractured horizontal well in unconventional reservoir, the early linear flow was expected to

163

Page 181: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

be dominant and could last for years depending on the formation permeability [17]. It is shown in

Figure 2.17 in Chapter 2 [16].

Chen and Raghavan (1997) proposed the third flow regime was early radial flow as shown in

Figure 2.18 in Chapter 2 [15]. This flow regime happened when the fluid flow from the fracture

tip within the fracture toward the horizontal wellbore [15]. This flow regime depended on the

fracture length and spacing [15]. It happened after the early linear flow and before the fracture

interference [15]. It was only observed when the fracture was very short or far apart [15]. Chen

and Raghavan (1997) proposed the fourth flow regime was compound linear flow as shown in

Figure 2.19 in Chapter 2 [15, 16]. It occurred once the fractures had interfered with each other.

The fluid flowed from the unstimulated rock volume toward the stimulated rock volume. Chen and

Raghavan (1997) proposed the fifth flow regime is late radial flow as shown in Figure 2.20 in

Chapter 2 [15, 16]. The flow occurred around the multifractured horizontal well boundaries [15].

The flow pattern was similar to the late time production of the vertically fractured well. It only

occurred if the well existed all alone in an undeveloped field and usually it required very long

production time and area to be developed in tight unconventional reservoir [15, 16]. Chen and

Raghavan (1997) proposed the boundary dominated flow [15]. This flow could be either a pseudo

steady state flow (no flow boundaries) or steady state flow (constant pressure boundaries) [15].

5.2.4 Flow Regions of Multi-Fractured Horizontal Well in Naturally

Fractured Reservoir

This section will present literature that discusses the flow regimes occurring in a multifractured

horizontal well. Ozkan et al. (2009) and Brown et al. (2009) proposed a trilinear flow model where

the drainage volume of multifractured horizontal well was limited to the inner reservoirs between

the fractures [10, 18]. It is shown in Figure 2.21 in Chapter 2. The basis of the trilinear flow model

assumed the production life of the multifractured horizontal well was dominated by the linear flow

regimes. The trilinear flow model coupled the linear flow in three adjacent flow regions. The flow

regions were the outer reservoir, the inner reservoir between fractures and the hydraulic fractures.

The uniform distribution of identical hydraulic fractures along the length of the horizontal well

was assumed.

164

Page 182: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

To allow production from the inner reservoir region between the hydraulic fractures, the region

is assumed to have natural fractures (dual porosity model). The flow regime in the inner reservoir

region is assumed to be transient flow and the model use the transient interporosity coefficient

(storativity ratio and interporosity flow coefficient).

The latest model is a horizontal well multifractured enhanced fracture region model from

Stalgorova and Mattar (2012) as shown in Figure 2.22 and Figure 2.23 in Chapter 2 [19]. This

model used the same concept with Ozkan et al. (2009) [18] but it assumed the unstimulated

reservoir region beyond the fracture tip contribution was negligible and the unstimulated reservoir

region between the fractures contribution was considered. Stalgorova and Mattar (2012) adopted

the branch fracture concept from Daneshy (2003) as shown in Figure 2.24 in Chapter 2 [20].

Daneshy (2003) explained that the branched fracturing could be caused by wellbore inclination

respect to in-situ stresses, perforation pattern and natural fractures. It can also be caused by low

anisotropy in-situ stresses [20].

5.3 Rate Transient Analysis (RTA) in Naturally Fractured

Reservoir

5.3.1 RTA Concept

There are different methods available to analyze production data. The two distinct methods are

type curve and non-type curve methods. Arps (1945) was the first to develop production data

analysis methods [21]. He developed decline curves for oil and gas production during transient

flow. The traditional decline analysis has limitations: it was not able to disassociate the production

forecast from operating conditions [21]. He assumed the historical operating condition stayed

constant for future production.

Fetkovich (1980) extended the decline curve concept into production data analysis where

before the type curve concept was used for pressure transient analysis [22]. He found that late time

(boundary dominated flow) data could be matched to type curves. The same methods from Arps

(1945) were used [81]. Both of these traditional decline curves relied on matching the model with

165

Page 183: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

the production data. The limitations were the assumptions of the productions parameter would

remain constants throughout time.

Recent methods such as Wattenbarger (1998), Blasingame et al. (1991), and Agarwal et al.

(1998) considered variable production parameters [14, 23, 24]. The improvement on traditional

analysis was the use of a normalized rate using the pressure drop (q/Δp) that allowed the effect of

pressure changes to be taken into account in the analysis. The pseudotime concept is the time

function for gas reservoirs that took into account compressibility changes of gas with pressure that

would allow the gas material balance to be dealt with carefully as the reservoir pressure decreased

with time. Blasingame et al. (1991) provided the typecurves for radial flow, elliptical well,

fractured vertical well, horizontal well with no fractures, finite conductivity fractures and infinite

conductivity fractures [23]. They used the same procedure from Wattenbarger et al. (1998) by

plotting the logarithm of the normalized rate with the logarithm of the material balance pseudo

time with another option of having rate integral and rate derivative on the y axis [23, 24]. Their

limitations were that the rate integral was very sensitive with early time errors and did not

distinguish the different flow regimes.

𝑡𝑡𝐶𝐶𝑡𝑡 =�𝜇𝜇𝑔𝑔𝑐𝑐𝑡𝑡�𝑖𝑖𝑞𝑞𝑔𝑔

�𝑞𝑞𝑔𝑔𝜇𝜇𝑔𝑔��� 𝑐𝑐𝑡𝑡�

𝑡𝑡

0

𝐼𝐼𝑡𝑡

where tca is material balance pseudo time. µg is gas viscosity, ct is total compressibility, qg is gas

flow rate, i is initial condition, and t is time.

Wattenbarger et al. (1998) typecurves were used to analyze linear flow specifically in tight

reservoirs where the linear flow could be dominant and last for years [24]. They assumed a vertical

well with fractures in the center of a rectangular reservoir where the fractures were assumed to

reach the reservoir boundaries. They used the log log plot of the normalized rate with material

balance pseudo time with another option of having pressure derivative on the y axis. Their

limitation was that the typecurve is only applicable for the linear flow and not for the boundary

dominated flow. Blasingame et al. (1991) provided the typecurves for radial flow and fractured

vertical well [23]. They used the same procedure as Wattenbarger et al. (1998) but with different

transient characterization using dimensionless reservoir boundaries parameters [23]. It was found

that this typecurve was more unique than Wattenbarger et al. (1998). But all these available

typecurves only applicable for vertical wells both fractured and unfractured.

166

Page 184: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5.4 Hoadley Field Properties

The Glauconitic Formation is a naturally fractured reservoir with natural fractures that are

identified from microseismic measurement and low cohesion values from Mohr-Coulomb failure

envelope. Type A or Type B reservoir storage would fit the description of the Glauconitic

Formation.

5.4.1 Production Data Review

Production history during eight months from November 1st, 2012 to July 20th, 2013 is used as

shown in Figure 5.1. There were time intervals that have zero gas flow rate; these were caused by

shut-in of the well to swab the well (unload the liquid within the wellbore). The initial gas flow

rate was 40 e3m3/d with average casing pressure of 3.5 MPa. The casing pressure (bottom hole

pressure or BHP) was chosen because of the continuity compared to the tubing pressure during

production.

Fig. 5.1: Production data history.

0

1

2

3

4

5

6

7

8

05

1015202530354045

0 25 50 75 100 125 150 175 200 225 250 275

BHP

(MPa

)

Gas P

rodu

ctio

n (e

3m3d

/d)

Production time (days)

Production 8 Months

Gas Production BHP

167

Page 185: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5.5 Methodology

5.5.1 Nonlinear Partial Differential Diffusivity Equation Solution

5.5.1.1 Nonlinear Partial Differential Diffusivity Equation For Real Gas

Derivation

A nonlinear partial differential diffusivity equation is derived for real gases by using the

continuity equation and Darcy’s law. The continuity equation and Darcy’s law are given by [25,

26]:

𝜕𝜕𝜕𝜕𝜕𝜕

(𝜌𝜌𝑣𝑣) = 𝜕𝜕𝜕𝜕𝑡𝑡

(𝜌𝜌𝜌𝜌) (6)

𝑣𝑣 = −0.006328 𝑘𝑘𝜇𝜇𝜕𝜕𝑃𝑃𝜕𝜕𝜕𝜕

(7)

𝑣𝑣 = −0.000264 𝑘𝑘𝜇𝜇𝜕𝜕𝑃𝑃𝜕𝜕𝜕𝜕

(8)

where ρ is fluid density, ν is fluid velocity, φ is porosity, k is permeability, µ is fluid viscosity, P

is fluid pressure, and x is distance. The unit conversion of 0.006328 is used to convert all the units

into ft3/d and Equation (8) is for flow rate in ft3/hours. Inserting Equation (8) into Equation (6)

produces:

𝜕𝜕𝜕𝜕𝜕𝜕�0.000264 𝑘𝑘

𝜇𝜇𝜕𝜕𝑃𝑃𝜕𝜕𝜕𝜕� = 𝜕𝜕

𝜕𝜕𝑡𝑡(𝜌𝜌𝜌𝜌) (9)

To develop the solution of Equation (9) for gas, additional equations are required:

𝜌𝜌𝑔𝑔 = 𝑀𝑀𝑆𝑆𝑅𝑅

𝑃𝑃𝑍𝑍

(10)

Inserting Equation (10) into Equation (9) yields:

𝜕𝜕𝜕𝜕𝜕𝜕�𝑀𝑀𝑆𝑆𝑅𝑅

𝑘𝑘𝑃𝑃𝜇𝜇𝑍𝑍

𝜕𝜕𝑃𝑃𝜕𝜕𝜕𝜕� = 1

0.000264𝜕𝜕𝜕𝜕𝑡𝑡�𝜌𝜌 𝑀𝑀

𝑆𝑆𝑅𝑅𝑃𝑃𝑍𝑍� (11)

Divide Equation (11) by M/RT produces:

𝜕𝜕𝜕𝜕𝜕𝜕�𝑘𝑘𝑃𝑃𝜇𝜇𝑍𝑍

𝜕𝜕𝑃𝑃𝜕𝜕𝜕𝜕� = 1

0.000264𝜕𝜕𝜕𝜕𝑡𝑡�𝜌𝜌 𝑃𝑃

𝑍𝑍� (12)

By using product rule for the right side of Equation (12) produces:

168

Page 186: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

𝜕𝜕𝜕𝜕𝑡𝑡�𝜌𝜌

𝑃𝑃𝑍𝑍� =

𝑃𝑃𝑍𝑍𝜕𝜕𝜌𝜌𝜕𝜕𝑡𝑡

+ 𝜌𝜌𝜕𝜕𝜕𝜕𝑡𝑡�𝑃𝑃𝑍𝑍�

𝜕𝜕𝜕𝜕𝑡𝑡�𝜌𝜌

𝑃𝑃𝑍𝑍� =

𝑃𝑃𝑍𝑍𝜕𝜕𝜌𝜌𝜕𝜕𝑃𝑃

𝜕𝜕𝑃𝑃𝜕𝜕𝑡𝑡

+ 𝜌𝜌𝜕𝜕𝜕𝜕𝑃𝑃

�𝑃𝑃𝑍𝑍�𝜕𝜕𝑃𝑃𝜕𝜕𝑡𝑡

𝜕𝜕𝜕𝜕𝑡𝑡�𝜌𝜌 𝑃𝑃

𝑍𝑍� = 𝜑𝜑𝑃𝑃

𝑍𝑍𝜕𝜕𝑃𝑃𝜕𝜕𝑡𝑡�1𝜑𝜑𝜕𝜕𝜑𝜑𝜕𝜕𝑃𝑃

+ 𝑍𝑍𝑃𝑃𝜕𝜕𝜕𝜕𝑃𝑃�𝑃𝑃𝑍𝑍�� (13)

Addition equations are needed such as:

𝑐𝑐𝑓𝑓 = 1𝜑𝜑𝜕𝜕𝜑𝜑𝜕𝜕𝑡𝑡

(14)

𝑐𝑐𝑔𝑔 = 1𝑃𝑃− 1

𝑍𝑍𝜕𝜕𝑃𝑃𝜕𝜕𝑍𝑍

= 𝑍𝑍𝑃𝑃𝜕𝜕𝜕𝜕𝑃𝑃�𝑃𝑃𝑍𝑍� (15)

If other fluids and formation compressibility are negligible the cg is assumed to be equal to ct.

Inserting Equations (13), (14) and (15) into Equation (12) produces:

𝜕𝜕𝜕𝜕𝜕𝜕�𝑘𝑘𝑃𝑃𝜇𝜇𝑍𝑍

𝜕𝜕𝑃𝑃𝜕𝜕𝜕𝜕� = 1

0.000264𝜌𝜌𝜇𝜇𝑐𝑐𝑡𝑡

𝑃𝑃𝜇𝜇𝑍𝑍

𝜕𝜕𝑃𝑃𝜕𝜕𝑡𝑡

(16)

A simplification from Equation (17) is used into Equation (16) produces Equation (18):

𝑃𝑃𝐼𝐼𝑃𝑃 = 12𝐼𝐼(𝑃𝑃2) (17)

𝜕𝜕𝜕𝜕𝜕𝜕� 𝑘𝑘𝑃𝑃2𝜇𝜇𝑍𝑍

𝜕𝜕𝜕𝜕𝜕𝜕

(𝑃𝑃2)� = 10.000264

𝜌𝜌𝜇𝜇𝑐𝑐𝑡𝑡1

2𝜇𝜇𝑍𝑍𝜕𝜕𝜕𝜕𝑡𝑡

(𝑃𝑃2) (18)

Divide both sides on Equation (18) by 2 produces:

𝜕𝜕𝜕𝜕𝜕𝜕�𝑘𝑘𝑃𝑃𝜇𝜇𝑍𝑍

𝜕𝜕𝜕𝜕𝜕𝜕

(𝑃𝑃2)� = 10.000264

𝜌𝜌𝜇𝜇𝑐𝑐𝑡𝑡1𝜇𝜇𝑍𝑍

𝜕𝜕𝜕𝜕𝑡𝑡

(𝑃𝑃2) (19)

Equation (19) can be expanded to 2D as follows:

𝜕𝜕𝜕𝜕𝜕𝜕�𝑘𝑘𝑥𝑥𝜇𝜇𝑍𝑍

𝜕𝜕(𝑃𝑃2)𝜕𝜕𝜕𝜕

� + 𝜕𝜕𝜕𝜕𝜕𝜕�𝑘𝑘𝑦𝑦𝜇𝜇𝑍𝑍

𝜕𝜕(𝑃𝑃2)𝜕𝜕𝜕𝜕

� = 10.000264

𝜑𝜑𝜇𝜇𝐶𝐶𝑡𝑡𝜇𝜇𝑍𝑍

𝜕𝜕(𝑃𝑃2)𝜕𝜕𝑡𝑡

(20)

169

Page 187: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5.5.1.2 Matlab Nonlinear PDE Toolbox and Code Editor

Equation (18) is a nonlinear parabolic PDE diffusivity equation. Matlab has a nonlinear

parabolic PDE toolbox (also referred to a PDE nonlinear code editor). The Matlab PDE nonlinear

code editor and command steps are explained below. Based on nonlinear parabolic diffusivity PDE

for gas derivation in the previous section. It is concluded to use the following simplification:

𝑢𝑢 = 𝑃𝑃2

𝜇𝜇𝑍𝑍 (21)

𝑃𝑃 = �𝑢𝑢𝜇𝜇𝑍𝑍 (22)

𝑑𝑑𝑃𝑃𝑑𝑑𝜕𝜕

= �𝜇𝜇𝑍𝑍 12𝑢𝑢−

12𝑑𝑑𝑢𝑢𝑑𝑑𝜕𝜕

(23)

𝑑𝑑𝑃𝑃𝑑𝑑𝜕𝜕

= �𝜇𝜇𝑍𝑍 12𝑢𝑢−

12𝑑𝑑𝑢𝑢𝑑𝑑𝜕𝜕

(24)

Equation (18) becomes:

𝜕𝜕𝜕𝜕𝜕𝜕�𝑘𝑘𝜕𝜕 𝜕𝜕𝑢𝑢

𝜕𝜕𝜕𝜕� + 𝜕𝜕

𝜕𝜕𝜕𝜕�𝑘𝑘𝜕𝜕 𝜕𝜕𝑢𝑢

𝜕𝜕𝜕𝜕� = 𝜑𝜑𝜇𝜇𝐶𝐶𝑡𝑡

0.000264𝜕𝜕𝑢𝑢𝜕𝜕𝑡𝑡

(25)

The porosity as a function of pressure is defined below by Bear et al. (2012) [25]:

𝜌𝜌 = 𝜌𝜌𝑖𝑖�1 − 𝑐𝑐𝑡𝑡(𝑃𝑃0 − 𝑃𝑃)� (26)

The permeability as a function of porosity and specific surface is based on the Kozeny-Carman

equation is defined below by Carman (1956), Timur (1968), and Gates (2011)) [26, 27, 28]:

𝑘𝑘 = 𝜑𝜑3

5𝑆𝑆2(1−𝜑𝜑)2 (27)

Equation (27) has the permeability term in m2, in term of mD, Equation (27) becomes:

𝑘𝑘 = 𝜑𝜑3

5𝑆𝑆2(1−𝜑𝜑)2 1015 (28)

Rearranging permeability as a function of pressure yields:

𝑘𝑘 = 𝜑𝜑𝑖𝑖�1−𝐶𝐶𝑡𝑡(𝑃𝑃0−𝑃𝑃)� 3

5𝑆𝑆2�1−�𝜑𝜑𝑖𝑖�1−𝐶𝐶𝑡𝑡(𝑃𝑃0−𝑃𝑃)� ��2 1015 (29)

170

Page 188: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The specific surface in (m2/m3) can be determined by using the initial permeability and initial

porosity (the initial permeability and porosity are determined from Chapter 3 as the final

permeability and porosity at the end of hydraulic fracturing). The permeability term in Equation

(29) needs to be defined in terms of u to solve the problem; this gives:

𝑘𝑘 =𝜑𝜑𝑖𝑖�1−𝐶𝐶𝑡𝑡��𝑢𝑢0𝜇𝜇𝑍𝑍 −�𝑢𝑢𝜇𝜇𝑍𝑍 �� 3

5𝑆𝑆2�1−�𝜑𝜑𝑖𝑖�1−𝐶𝐶𝑡𝑡��𝑢𝑢0𝜇𝜇𝑍𝑍−�𝑢𝑢𝜇𝜇𝑍𝑍�� ��2 1015 (30)

Equations (23), (24), and (25) are used in the following section on the Matlab PDE nonlinear

code editor/command to solve the problem. To simplify the analysis, gas viscosity and gas

compressibility factor are both assumed constant. The values used in the calculations are the

average values of gas viscosity and gas compressibility factor between the initial pressure (20

MPa) and wellbore pressure (3.449 MPa).

First, the problem parameters need to be defined. The gas viscosity, reservoir initial porosity,

reservoir total compressibility (fracture compressibility is assumed to be much higher than

formation and gas compressibility), and Kozeny-Carman specific surface values are defined.

Table 5.1: Input parameters.

Average gas viscosity (cP) 0.02

Average gas Z-factor 0.78

Initial SRV porosity (fraction) 0.16

Fracture compressibility (1/psi) 4.02E-04

Kozeny-Carman specific surface (m2/m3) 7044

Second, the geometry and mesh need to be defined. The SRV and wellbore geometry are

defined in a rectangular region (x is horizontal axis and y is vertical axis). The region is discretized

using triangular mesh elements where the maximum element size is defined in Chapter 3. The

geometry and the mesh are shown in Figures 5.2 to 5.3 and Table 5.2. The mesh was generated

by using the Matlab PDE toolbox.

171

Page 189: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.2: Geometry with edge labels displayed.

172

Page 190: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.3: SRV with triangular element mesh.

Table 5.2: Matlab PDE nonlinear geometry and mesh.

SRV length (m) in Y coordinate 174

SRV width (m) in X coordinate 60

Wellbore production port length (m) 1

Wellbore production width length (m) 1

Triangular element maximum size (m) 5

Third, the boundary conditions are defined. The pressure boundary conditions are defined for

wellbore conditions using wellbore pressure during production from the field data. The u term for

173

Page 191: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

the wellbore pressure boundary conditions is calculated by using wellbore pressure, average gas

viscosity, and average gas compressibility factor, listed in Table 5.3.

Table 5.3: Matlab PDE nonlinear boundary conditions.

Wellbore pressure (psi) 500

u at wellbore pressure (psi2/cP) 16224497

Fourth, the coefficients of the nonlinear PDE as used by the Matlab toolbox are defined. Matlab

defines the parabolic nonlinear PDE as a function of a, c, d, and f coefficients:

𝐼𝐼 𝜕𝜕𝑢𝑢𝜕𝜕𝑡𝑡− ∇(𝑐𝑐∇𝑢𝑢) + 𝑡𝑡𝑢𝑢 = 𝑓𝑓 (31)

Based on Equation (31) the parabolic nonlinear PDE coefficients are:

𝑐𝑐 = 𝑘𝑘 =𝜑𝜑𝑖𝑖�1−𝐶𝐶𝑡𝑡��𝑢𝑢0𝜇𝜇𝑍𝑍 −�𝑢𝑢𝜇𝜇𝑍𝑍 �� 3

5𝑆𝑆2�1−�𝜑𝜑𝑖𝑖�1−𝐶𝐶𝑡𝑡��𝑢𝑢0𝜇𝜇𝑍𝑍−�𝑢𝑢𝜇𝜇𝑍𝑍�� ��2 1010 (32)

Equation (32) used unit conversion 1010 to convert 105 ft3/hours (field flow rate is in magnitude

of 105 ft3/days) to ft3/hours.

𝐼𝐼 = 𝜑𝜑𝜇𝜇𝐶𝐶𝑡𝑡0.000264

= 𝜑𝜑𝑖𝑖�1−𝐶𝐶𝑡𝑡(𝑃𝑃0−𝑃𝑃)� 𝜇𝜇𝐶𝐶𝑡𝑡0.000264

(33)

𝑡𝑡 = 0 (34)

𝑓𝑓 = 0 (35)

The parabolic nonlinear PDE is solved using the wellbore production time (8 months or 6,300

hours) as shown in Table 5.4. Then the parabolic nonlinear PDE is solved using the Matlab code

editor lines as a function of u0, production time, boundary condition, and PDE nonlinear

coefficients. The detailed Matlab nonlinear parabolic PDE code is listed in Appendix A.

174

Page 192: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 5.4: Matlab PDE nonlinear simulation time.

Time (hours) 6300

Time simulation spacing (hours) 157.5

Gas flow rate is calculated using Darcy’s law and Matlab simulation outputs (permeability and

pressure drop gradient). The simulated gas flow rate is then compared with the field gas flow rate.

5.5.1.3 Matlab Nonlinear PDE Code Editor Results

The Matlab nonlinear PDE solver uses the Galerkin finite element method to solve the PDE

together with forward Euler time integrator and Newton’s method to deal with the nonlinearity of

the PDE. The default relative tolerance of 1x10-3 and absolute tolerance of 1x10-4 are used in the

Matlab nonlinear PDE solver. The triangular mesh is created for the 2D geometry using Delaunay

triangulation. The mesh size is determined from the geometry shape and the defined maximum

mesh size of 5 m.

After u is found, then the pressure, porosity, and permeability are calculated by using Equations

(22), (26), and (29). The results from Matlab is presented in contour plots at three different

production times: early production time (t=157.5 hours), mid production time (t=2,992.5 hours),

and final production time (t=6,300 hours). The pressure plots are presented in Figures 5.4a to 5.4c,

permeability in Figures 5.5a to 5.5c, and porosity in Figures 5.6a to 5.6c. The pressure and porosity

contour plots show small changes during early, mid, and late production time. But the permeability

contour plots show great changes during early, mid, and late production time.

175

Page 193: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

(c)

Fig. 5.4: Pressure contour plot in SRV during (a) 157.5 hours, (b) 2,992.5 hours, and (c) 6,300 hours production.

176

Page 194: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

(c)

Fig. 5.5: Permeability contour plot in SRV during (a) 157.5 hours, (b) 2,992.5 hours, and (c) 6,300 hours production.

177

Page 195: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

(c)

Fig. 5.6: Porosity contour plot in SRV during (a) 157.5 hours, (b) 2,992.5 hours, and (c) 6,300 hours production.

178

Page 196: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The pressure plots are presented in Figures 5.7a and 5.7b, permeability in Figures 5.8a and

5.8b, and porosities in Figure 5.9a and 5.9b. From Figure 5.7b, the pressure of 11.8 MPa results at

two meters distance from wellbore during the end of production (262.5 days). This pressure is

higher than the initial reservoir pressure (9 MPa). The pressure of 11.8 MPa produces permeability

of 2,763 mD. These responses show that the SRV pressure and permeability are still high and

allow the SRV to be drained for later production.

179

Page 197: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 5.7: Pressure profile as a function of (a) time and (b) distance from SRV boundary to wellbore.

0

5

10

15

20

25

0 50 100 150 200 250 300

Pres

sure

, MPa

Time, days

Pressure During 8 Months Production within SRV

y=18.5 m y=32 m y=45.6 m y=48.2 m y=55.4 my=59.5 m y=60.8 m y=64 m y=67 m y=70.5 my=70.7 m y=74.6 m y=78.6 m y=80.1 m y=81 my=85 m y=87 m

0

5

10

15

20

25

0 20 40 60 80 100

Pres

sure

, MPa

Distance, m

Pressure from SRV Boundary to Wellbore

t=0 days t=0.4 days t=0.8 days t=1.3 dayst=1.7 days t=2.1 days t=6.6 days t=13.1 dayst=19.7 days t=26.3 days t=32.8 days t=39.4 dayst=46 days t=52.5 days t=59.1 days t=65.6 dayst=72.2 days t=78.8 days t=85.3 days t=92 dayst=98.4 days t=105 days t=111.6 days t=118 dayst=124.7 days t=131.3 days t=137.8 days t=144.4 dayst=151 days t=157.5 days t=164 days t=170.6 dayst=177.2 days t=183.8 days t=190.3 days t=197 dayst=203.4 days t=210 days t=217 days t=223 dayst=230 days t=236.3 days t=243 days t=249.4 dayst=256 days t=262.5 days

180

Page 198: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 5.8: Permeability profile as a function of (a) time and (b) distance from SRV boundary to wellbore.

0

5,000

10,000

15,000

20,000

25,000

30,000

0 50 100 150 200 250 300

Perm

eabi

lity,

mD

Time, days

Permeability During 8 Months Production within SRV

y=32 m y=45.6 m y=48.2 m y=55.4 my=59.5 m y=60.8 m y=64 m y=67 my=70.5 m y=70.7 m y=74.6 m y=78.6 my=80.1 m y=81 m y=85 m y=87 m

0

5,000

10,000

15,000

20,000

25,000

0 20 40 60 80 100

Perm

eabi

lity,

mD

Distance, m

Permeability from SRV Boundary to Wellbore

t=0 days t=0.4 days t=0.8 days t=1.3 dayst=1.7 days t=2.1 days t=6.6 days t=13.1 dayst=19.7 days t=26.3 days t=32.8 days t=39.4 dayst=46 days t=52.5 days t=59.1 days t=65.6 dayst=72.2 days t=78.8 days t=85.3 days t=92 dayst=98.4 days t=105 days t=111.6 days t=118 dayst=124.7 days t=131.3 days t=137.8 days t=144.4 dayst=151 days t=157.5 days t=164 days t=170.6 dayst=177.2 days t=183.8 days t=190.3 days t=197 dayst=203.4 days t=210 days t=217 days t=223 dayst=230 days t=236.3 days t=243 days t=249.4 dayst=256 days t=262.5 days

181

Page 199: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 5.9: Porosity profile as a function of (a) time and (b) distance from SRV boundary to wellbore.

0.000.020.040.060.080.100.120.140.160.18

0 50 100 150 200 250 300

Poro

sity,

frac

tion

Time, days

Porosity During 8 Months Production within SRV

y=18.5 m y=32 m y=45.6 m y=48.2 m y=55.4 my=59.5 m y=60.8 m y=64 m y=67 m y=70.5 my=70.7 m y=74.6 m y=78.6 m y=80.1 m y=81 my=85 m y=87 m

0.000.020.040.060.080.100.120.140.160.18

0 20 40 60 80 100

Poro

sity,

frac

tion

Distance, m

Porosity from SRV Boundary to Wellbore

t=0 days t=0.4 days t=0.8 days t=1.3 dayst=1.7 days t=2.1 days t=6.6 days t=13.1 dayst=19.7 days t=26.3 days t=32.8 days t=39.4 dayst=46 days t=52.5 days t=59.1 days t=65.6 dayst=72.2 days t=78.8 days t=85.3 days t=92 dayst=98.4 days t=105 days t=111.6 days t=118 dayst=124.7 days t=131.3 days t=137.8 days t=144.4 dayst=151 days t=157.5 days t=164 days t=170.6 dayst=177.2 days t=183.8 days t=190.3 days t=197 dayst=203.4 days t=210 days t=217 days t=223 dayst=230 days t=236.3 days t=243 days t=249.4 days

182

Page 200: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

The gas flow rate is calculated by using Darcy’s law and the permeability and pressure gradient

at the wellbore. This calculated gas flow rate is compared with the field gas flow rate. The initial

gas flow rate is matched by finding the total compressibility (field measured formation

compressibility of 3x10-6 1/psi and total compressibility of 5.81x10-4 1/psi). The matched total

compressibility of 4.0559x10-4 1/psi is similar to the total compressibility of 5.81x10-4 1/psi. This

total compressibility is assumed to be driven by the fracture compressibility (fluid and formation

compressibility are assumed to be small and neglected). The matched fracture compressibility is

compared to fracture compressibility from empirical correlations [6]. Aguilera (1999) presented

for net stress on fracture of 3,742 psi (net stress is reduction of total overburden stress of 6,642 psi

with pressure within SRV at the end of hydraulic fracture of 2,900 psi), the fracture compressibility

is between 3x10-5 1/psi (for 50% mineralization within fracture) and 1.25x10-4 1/psi (for zero

mineralization within fracture) [6]. The matched fracture compressibility is within the magnitude

of Aguilera’s (1999) fracture compressibility with zero mineralization within the fracture.

The cross section area on the production port considers the z-direction. A permeability of 0.3

mD at the wellbore tip is determined from the Matlab simulation and it is constant during

production. The simulated gas flow rate has overall good match with the field gas flow rate

especially from Day 25 to the end of the time period as shown in Figure 5.10. There are some

differences in early production time (smaller than 25 days) that may be caused by different pressure

drop gradient in the simulation and the field. A higher initial porosity of 0.17 is simulated, but the

simulated gas flow rate is similar (1 e3m3/day of differences) with the initial porosity of 0.16. This

concludes that the initial porosity does not affect the flow rate. The matched gas flow rate profile

shows that the model is applicable for the other fields with similar reservoir properties to the

Hoadley field.

183

Page 201: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

(a)

(b)

Fig. 5.10: Comparison of (a) field gas flow rate and (b) bottomhole pressure with new diffusion model with porosity of 0.16 and 0.17 (no changes).

0

10

20

30

40

50

0 50 100 150 200 250 300

Gas P

rodu

ctio

n, e

3m3/

d

Production time, days

Gas Production

q por=0.16 q por=0.17 q field

0

1

2

3

4

5

0 50 100 150 200 250 300

BHP,

MPa

Production time, days

Bottomhole Pressure

BHP Matlab BHP Field

184

Page 202: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5.5.2 Application of Rate Transient Analysis Simulator (IHS Harmony

Rate)

5.5.2.1 IHS Harmony Rate Transient Analysis Inputs

IHS Harmony was used to conduct rate transient analysis on an eight month production data

time interval for Well 1-1843-2W5 on the Hoadley Field after the hydraulic fracturing. The inputs

used for the analysis are listed in Table 5.5:

Table 5.5: IHS Harmony simulation input.

Entity name COPRC ET AL 100 HZ

WROSES 1-18-43-2

Primary fluid Gas

Country/Province Canada/Alberta

Field Hoadley

Formation Glauconitic

Initial gas reservoir P (kPa) 9188

Initial gas reservoir T (C) 70

Formation thickness (m) 43

Wellbore diameter (m) 0.025

Initial gas saturation (%) 50

Initial oil saturation (%) 15

Initial water saturation (%) 35

Total compressibility (1/psi) 5.81E-04

Formation compressibility (1/psi) 3.00E-06

Gas Z-factor 0.712

CO2 (%) 2.18

H2S (%) 0

N2 (%) 5.11

185

Page 203: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5.5.2.2 Identifying Flow Regimes Using Type Curves

All type curves identified that the flow regime was linear flow regime (transient flow regime)

and it had not reached the boundary dominated flow regime yet. The linear flow was from the

reservoir into the induced hydraulic fracture or activated pre-existing natural fractures. The type

curves could not be used to determine the reservoir parameters since the type curves were built for

fractured and un-fractured vertical wells. The median filter was applied to all available type curves

to provide better interpretation. The available type curves used were the Blasingame finite

conductivity fracture (Figure 5.11), Blasingame elliptical (Figure 5.12), Blasingame horizontal

(Figure 5.13), and Wattenbarger (Figure 5.14). Doublet et al. (1994) found the Blasingame type

curve to be useful in the modeling of elliptical flow that transitioned into boundary dominated flow

[29]. The Blasingame finite conductivity fracture represented a square or cylindrical reservoir with

finite conductivity fracture in the center, the Blasingame elliptical represented an elliptical

reservoir with a finite conductivity fracture in the center, and the Blasingame horizontal

represented a square reservoir with a horizontal well in the center. The Wattenbarger type curve

was also used to fit the field data. The type curve analysis was useful to analyze unconventional

gas reservoir with long term linear flow. Several points at Wattenbarger normalized rate plot and

derivative (half slope of derivative blue points) plot fitted nicely with the field data. The plots

showed the flow regime was linear flow regime. The Blasingame finite conductivity fracture type

curve fitted better with the field data and the type curve outputs were used for the unconventional

reservoir analysis inputs. The Blasingame finite conductivity fracture type curve fitted better with

the field data because fracture permeability of 23,400 mD and fracture aperture of 0.0047 ft (1.43

mm) from Chapter 3 analysis produced the fracture flow capacity of 109.98 mD.ft. Fracture flow

capacity of less than 10,000 mD.ft represented a finite conductivity fracture and the fracture flow

capacity was the product of fracture permeability and fracture aperture.

186

Page 204: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.11: Blasingame finite conductivity fracture with median filter.

187

Page 205: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.12: Blasingame elliptical with median filter.

188

Page 206: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.13: Blasingame horizontal with median filter.

189

Page 207: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.14: Wattenbarger with median filter.

5.5.2.3 Unconventional Reservoir Analysis (Unconventional Gas Module)

The input for the unconventional gas module were the reservoir properties, Blasingame finite

conductivity fracture type curves, number of hydraulic fracture stages (12), and effective

horizontal well length (hydraulic fracture stages port 1 to 12 distance of 1268.1 m), then the

simulator calculated the reservoir permeability.

The first plot was the square root of time plot showing a straight line which was confirming

the linear flow regime from the type curve with no departure from the straight line (boundary

dominated flow) (Figure 5.15a). The dotted green line (boundary dominated flow starting point)

was chosen at the end of the linear flow to get the optimistic assessment of the SRV that was

indicating the minimum SRV. The dotted green line was tied to the extrapolation red line on the

flowing material balance at Figure 5.15. The next plot is the log-log type curve between flow rate

and time representing the model from the square root of time plot as shown in Figure 5.16. The

red type curve showed a bounded drainage volume and the brown type curve assumed the linear

flow occurred for infinite time and it was not bounded by any drainage volume. The red type curve

190

Page 208: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

was an optimistic forecast for the well. Changing the slope at the log-log type curve would change

the skin values. The skin value that matched the type curve slope was zero (it was a hydraulically

fractured well). The input for this unconventional gas module was the horizontal wellbore length

and number of hydraulic fracture stage, then the software calculated the reservoir permeability

using the square root of time plot slope relationship with reservoir permeability.

Fig. 5.15: Unconventional gas module square root time plot to identify the pessimistic boundary dominated flow (green vertical line),

191

Page 209: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.16: Unconventional gas module typecurve plot to identify the pessimistic boundary dominated flow (green vertical line).

5.5.2.4 History Match Using Horizontal Multifractured Enhanced Fracture

Region Analytical Model

5.5.2.4.1 Case 1 With Constant FCD, SRV Half-Length, and Matrix

Permeability

To determine the SRV permeability and reservoir permeability, the reservoir properties and

the unconventional gas module output were applied into the horizontal multifractured enhanced

fracture region analytical model [19]. This model was chosen due to the assumption that there was

unstimulated reservoir volume between SRV from created hydraulic fractures. The input for this

analytical model were the reservoir properties and unconventional gas module simulation results.

These inputs were used as the simulation starting point. The horizontal multifractured enhanced

fracture region analytical model was used to simulate SRV permeability and reservoir

192

Page 210: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

permeability. The analytical model produced a history match between the gas flow rate and

flowing field data with the simulated data. To improve the history match, two cases were

simulated. Case 1 used constant dimensionless fracture conductivity (FCD) of 105, SRV half-

length of 6.7 m, and reservoir permeability (k2) of 0.00648 mD. Case 1 produced SRV

permeability (k1) of 827 mD (PSS model), 9,736 mD (slabs model), 1,136 mD (cubes model), and

15,516 mD (sticks model). The PSS model assumes the flow between matrix and fractures is in

pseudo steady state [16]. The transient model assumes the flow between matrix and fractures is

transient and it has three types of matrix geometry (slabs, cubes and sticks geometries) [16].The

results showed the transient models have a higher permeability compared to the PSS model which

might be caused by the transient models assuming the flow has only reached the closest SRV but

it has not reached the furthest SRV yet. Also the PSS model assumes the flow has already reached

all the no flow SRV boundaries. The results are listed on Table 5.6 with each model history match

shown in Figure 5.17 to 5.20. Case 1 produced a good match on the gas flow rate but only average

good match on the flowing pressure.

Case 1 produced a higher FCD compared to the expected FCD of 5.5. It also produced smaller

reservoir permeability (k2) compared to field reservoir permeability of 0.07 mD, smaller SRV half-

length compared to field SRV half-length of 87 m, a much smaller SRV permeability (k1)

compared to the Matlab results, and much smaller reservoir boundaries of around 23 m compared

to the field SRV half-length.

193

Page 211: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Table 5.6: Case 1 simulation results for constant FCD, SRV half-length and matrix permeability.

Simulation results Hz Multifrac Enhanced fracture region model

Dual porosity model

PSS slabs cubes sticks

FCD 105 105 105 105 k1 (mD) SRV 827 9736 1136 15516 k2 (mD) reservoir 0.0065 0.0065 0.0065 0.0065 xf (m) 6.7 6.7 6.7 6.7 ye (m) 23.8 23.7 23.8 23.7 omega 0.10 0.01 0.15 0.01 lamda 0.00 0.00 0.00 0.00 S 0.00 0.03 1.73 1.56 P Match Average good Average good Average good Average good Q Match Good Good Good Good

Fig. 5.17: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with PSS dual porosity model.

194

Page 212: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.18: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with slabs model.

195

Page 213: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.19: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with cubes model.

196

Page 214: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.20: History match with horizontal multifrac enhanced fracture region using unconventional reservoir analysis for constant FCD, SRV half-length and matrix permeability with sticks model.

5.5.2.4.2 Case 2 With Constant SRV Half-Length

Case 2 used a constant SRV half-length of 87 m (the SRV half-length is from Chapter 3) and

matrix permeability of 6.5 x10-5 mD. This case produced different values of FCD of 5 (PSS model),

2 (slabs model), 2 (cubes model), and 0.23 (sticks model). The simulated SRV permeability (k1)

were 13,528 mD (PSS model), 21,037 mD (slabs model), 22,288 mD (cubes model), and 22,345

mD (sticks model). The simulated matrix permeability (k2) were 6.5 x10-5 mD (PSS model),

6.5x10-5 mD (slabs model), 6.5x10-5 mD (cubes model), and 6.7x10-5 mD (sticks model). This case

produced smaller matrix permeability compared to the field data of 0.07 mD but it used the SRV

half-length of 87 m from the field data (microseismic). This case produced a similar FCD

compared to expected FCD of 5.5, much smaller matrix permeability (k2), a good range of SRV

permeability (k1) compared to the Matlab results (higher range compared to Case 1), and good

reservoir boundaries of around 200 m (longer than the field SRV half-length). In conclusion, Case

197

Page 215: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

2 is better compared to Case 1 due to a better match to the Matlab results and the field reservoir

properties except the matrix permeability. The smaller matrix permeability in the order of 6x10-5

mD could be expected in tight gas reservoir. The results are listed in Table 5.7 and Figure 5.21 to

5.24.

Table 5.7: Case 2 simulation results for constant SRV half-length.

Simulation results Hz Multifrac Enhanced fracture region model

Dual porosity model PSS slabs cubes sticks FCD 5 2 2 0.23 k1 (mD) SRV 13528 21037 22288 22345 k2 (mD) Matrix 6.5E-05 6.5E-05 6.5E-05 6.7E-05 xf (m) 87 87 87 87 ye (m) 249 218 205 198 omega 2.8E-02 3.4E-03 3E-03 2E-01 lamda 1.4E-08 3.8E-08 1.3E-07 2E-08 S 0.00 0.01 0.01 0.00 P Match Average good Average good Average good Average good Q Match Good Good Good Good

198

Page 216: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.21: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with PSS model.

199

Page 217: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.22: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with slabs model.

200

Page 218: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.23: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with cubes model.

201

Page 219: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Fig. 5.244: History match with horizontal multifrac enhanced fracture region for constant SRV half-length with sticks model.

5.6 Conclusions

A nonlinear diffusivity PDE is derived for the flow and pressure in the SRV and solved by

using Matlab PDE nonlinear code editor to determine the pressure, permeability, and porosity

profile as a function of distance and time within the SRV during the production. The porosity and

permeability are simulated as a function of the PDE solution (pressure). The simulated

permeability and pressure drop are used to simulate the gas flow rate. The conclusions of this

chapter are:

1. The production data could be reasonably well matched by using the new nonlinear diffusivity

theory. This match is done by finding appropriate total compressibility.

2. There are some gas flow rate differences with the field data at early time (simulated gas flow

rate is smaller than field gas flow rate). This may be caused by higher field pressure drop at

early time compared to simulated pressure drop (flow rate is affected by pressure drop). The

higher field pressure drop could be caused by proppant crushing due to fractures close and in-

202

Page 220: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

situ stress changes during early time production; factors which are not considered within PDE

diffusivity solution.

3. The matched gas flow rate profile shows that the model is applicable for other fields with

similar reservoir properties to the Hoadley field.

4. IHS Harmony Case 2 produces similar SRV permeability to the nonlinear PDE and reservoir

boundary to field for all models (PSS, slabs, cubes, and sticks) using field SRV half-length.

5. The nonlinear PDE SRV permeability is a function of distance and time during production

whereas the IHS Harmony SRV permeability is a single value.

6. The nonlinear PDE solution provides a better and more detailed view of the SRV permeability

distribution compared to that of IHS Harmony.

203

Page 221: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5.7 References

[1] Bulnes, A. C., and Fitting, R. U. 1945. An Introductory Discussion of the Reservoir Performance of Limestone Formations. Trans., AIME. Vol. 160, 179.

[2] Imbt, W. C., and Ellison, S. P. 1946. API Drill and Prod. Prac. 364. [3] Warren, J. E., and Root, P. J. 1963. The Behavior of Naturally Fractured Reservoirs. SPEJ

426. [4] Stearns, D. W. 1982-1994. AAPG Fractured Reservoirs School Notes. Great Falls, Montana. [5] Nelson, R. 1985. Geologic Analysis of Naturally Fractured Reservoirs. Contributions in

Petroleum Geology and Engineering, Vol. 1, Gulf Publishing Co., Houston, Texas. [6] Aguilera, R. 1998. Geologic Aspects of Naturally Fractured reservoirs. The Leading Edge,

pp. 1667-1670, December. [7] Aguilera, R. 2003. Geologic and Engineering Aspects of Naturally Fractured Reservoirs.

CSEG Recorder, February. [8] McNaughton, D. A. and Garb, F. A. 1975. Finding and Evaluating Petroleum Accumulations

in Fractured Reservoir Rock. Exploration and Economics of the Petroleum Industry, v.13, Matthew Bender & Company Inc.

[9] Aguilera, R. 1995. Naturally Fractured Reservoirs, Tulsa: PennWell Books. p.521. [10] Brown, M., Ozkan, E., Raghavan, R., and Kazemi, H. 2009. Practical Solutions for Pressure

Transient Responses of Fractured Horizontal Wells in Unconventional Reservoirs. Paper SPE 125043 presented at the 2009 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, USA, 4-7 October.

[11] Kazemi, H. 1969. Pressure Transient Analysis of Naturally Fractured Reservoirs with Uniform Fracture Distributions. Soc. Pet. Eng. Jour. (Dec.) 451-461: Trans., AIME, Vol. 261.

[12] de Swan-O, A. 1976. Analytical Solutions for Determining Naturally Fractured Reservoir Properties by Well Testing. Soc. Pet. Eng. Jour. (June) 117-122, Trans., AIME, Vol. 261.

[13] Serra, K., Reynolds, A. C., and Raghavan, R. 1983. New Pressure Transient Analysis Methods for Naturally Fractured Reservoirs. Jour. Pet. Tech.:2271-2283.

[14] Wattenbarger, R. A., El-Banbi, A. H., Villegas, M. E., and Maggard, J. B. 1998. Production Analysis of Linear Flow into Fractured Tight Gas Wells. Paper SPE 39931 presented at the 1998 Rocky Mountain Regional/Low Permeability Reservoirs Symposium and Exhibition, Denver, USA, 5-8 April.

[15] Chen, C. C., and Raghavan, R. 1997. A Multiply-Fractured Horizontal Well in a Rectangular Drainage Region. SPE J. (2): 455-465. SPE 37072 PA.

[16] Fekete reference material, 2015, Dual Porosity. http://www.fekete.com/SAN/WebHelp/FeketeHarmony/Harmony_WebHelp/Content/HTML_Files/Reference_Material/General_Concepts/Dual_Porosity.htm, sited: May 1st 2015.

[17] Nobakht, M., Clarkson, C., and Kaviani, D., 2011, New Type Curves for Analyzing Horizontal Well with Multiple Fractures in Shale Gas Reservoirs. Paper CSUG/SPE 149397 presented at the Canadian Unconventional Resources Conference, Calgary, Alberta, Canada, 15-17 November.

204

Page 222: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

[18] Ozkan, E., Brown, M., Raghavan, R., and Kazemi, H. 2009. Comparison of Fractured Horizontal –Well Performance in Conventional and Unconventional Reservoirs. Paper SPE 121290 presented at the 2009 SPE Western Regional Meeting, San Jose, California, USA, 24-26 March.

[19] Stalgorova, E., and Mattar, L. 2012. Practical Analytical Model to Simulate Production of Horizontal Wells with Branch Fractures. Paper SPE 162515 presented at the SPE Canadian Unconventional Resources Conference, Calgary, Alberta, Canada, 30 October – 1 November.

[20] Daneshy, A. A. 2003. Off Balance Growth: A New Concept in Hydraulic Fracturing. Journal of Petroleum Technology 55(4): 78-85. SPE 80992-MS. http:dx.doi.org/10.2118/80992-MS.

[21] Arps, J. J. 1945. Analysis of Decline Curves. Trans. AIME, 160, 228. [22] Fetkovich, M. J. 1980. Decline Curve Analysis using Type Curves. JPT (June), 1065. [23] Blasingame, T. A., McGray, T. I., Lee, W. J. 1991. Decline Curve Analysis for Variable

Pressure Drop/Variable Flow Rate Systems. Paper SPE 21513 presented at the SPE Gas Technology Symposium, 23-24 January.

[24] Agarwal, R. G., Gardner, D. C., Kleinsteiber, S. W. and Fussell, D. D. 1998. Analyzing Well Production Data Using Combined Type Curve and Decline Curve Concepts. Paper SPE 57916 presented at the 1998 SPE Annual Technical Conference and Exhibition, New Orleans, 27-30 September.

[25] Batchelor, G. K. 1967. An Introduction to Fluid Dynamics. Cambridge: Cambridge Mathematical Library.

[26] Darcy, H. 1856. Les Fontaines Publiques de la Ville de Dijon, Dalmont, Paris. [27] Bear, J., Tsang, C-F., and De Marsily, G. 2012. Flow and Contaminant Transport in

Fractured Rock. San Diego: Academic Press Inc. [28] Carman, P. C. 1956. Flow of Gases Through Porous Media. New York City: Academic Press

Inc. [29] Timur, A. 1968. An Investigation of Permeability, Porosity, and Residual Water Saturation

Relationships for Sandstone Reservoirs. The Log Analyst 9 (4). [30] Gates, I. D. 2011. Basic Reservoir Engineering. Dubuque: Kendall Hunt Publishing

Company. [31] Doublet, L. E., Pande, P. K., McCollum, T. J., Blasingame, T. A. 1994. Decline Curve

Analysis Using Type Curves-Analysis of Oil Well Production Data Using Material Balance Time Application to Field Cases. Paper SPE 28688 was presented at the 1994 Petroleum Conference and Exhibition of Mexico, Veracruz, Mexico, 10-13 October. http://dx.doi.org.ezproxy.lib.ucalgary.ca/10.2118/28688-MS

205

Page 223: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

CHAPTER 6: CONCLUSIONS AND RECOMMENDATIONS

6.1 Conclusions

The main objective of this study was to optimize the hydraulic fracture design during multi-

stages hydraulic fracture in horizontal well of tight gas formations with pre-exiting natural

fractures by determining the stimulated rock volume (SRV) characteristics such as enhanced

permeability, fracture spacing, pressure and in-situ stress changes profile during hydraulic

fracturing and production. To meet the objective, the study was divided into three different parts.

In the first part, the impact of SRV dimensions and Young’s modulus on the SRV effective

permeability during hydraulic fracturing using a 3D finite element analysis (Abaqus FEA) was

evaluated. The first part also included an investigation of fracture aperture, spacing, and number

within SRV using a new semi-analytical approach.

The second part investigated the impact of rock mechanical properties and injected volume

during hydraulic fracturing on SRV dimensions using a new analytical model. The model was

calibrated using microseismic data.

The third part explored a nonlinear diffusivity PDE solution using Matlab PDE nonlinear and

IHS Harmony to analyze field gas flow rate and pressure. The third part solved SRV permeability,

pressure, and porosity as a function of distance and time, and gas flow rate with field data.

The novel aspects of the thesis are as follows. First, for the first time, the developed workflow

using the 3D finite element analysis and the semi-analytical approach to characterize the fracture

network characteristics within the SRV during hydraulic fracturing was applied for the Glauconitic

Formation (Hoadley, AB). Second, this study establishes a new analytical SRV dimensions model

that combines rock mechanical properties and injected volume during hydraulic fracturing and

uses microseismic monitoring data for calibration. Third, for the first time the nonlinear diffusivity

PDE solution was solved to determine the SRV permeability, pressure, and porosity as a function

of distance and time and then the solutions are compared with the RTA simulator (IHS Harmony)

results.

206

Page 224: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

From this research, several conclusions have been made:

1. Finite element analysis shows that a reduction of Young’s modulus decreases the effective

permeability. A reduction of SRV dimensions by 10% increases the effective permeability

within SRV and decreases the pressure drop along SRV length.

2. A hydraulic fracture induces an increase in the total in-situ stress values. Changes in the total

in-situ stresses calculated by the finite element analysis are overestimated. The overestimated

in-situ stresses result from the unavailability of core testing on Biot’s constant and the Biot’s

constant is assumed equal to unity. This implies that the finite element analysis calculates the

maximum impact of hydraulic fracturing on the increase of in-situ stresses. It also implies that

the total in-situ stresses increase in the field induced by hydraulic fracturing is lower than the

in-situ stress increase calculated by the finite element analysis.

3. The new semi-analytical approach results reveal that a reduction of Young’s modulus

decreases the number of major fractures for constant fracture aperture. This implies that the

harder the rock, it is easier to be hydraulically fractured.

4. SRV growth is not apparent for some of hydraulic fracture stages. In some stages, microseismic

events are not triggered in the proximity of horizontal wellbore due to effect of depleted zones

surrounding old wellbores near the horizontal wellbore. This implies that the hydraulic fracture

fluid may grow toward neighboring wells and increase the pressure within the depleted zone

without creating shear failure and consequent microseismic events.

5. Some SRV dimensions are shorter than interpreted SRV dimensions via microseismic due to

the presence of depleted zones around old wellbores. During hydraulic fracturing, the injection

fluid reopens natural fractures in proximity of the depleted zone which triggers microseismic

events. The results suggest that these reopened natural fractures do not contribute to the gas

production.

6. The changes of the total stress and their impact on SRV dimensions reveal that total stresses

must be included to avoid overestimation of the dimensions of the SRV.

7. The nonlinear PDE model derived for the SRV yields results that are consistent with the field

gas production profiles. The only history match parameter adjusted to match the field data was

the total compressibility. The theory provides a starting point for multidimensional methods to

generate type curves for tight gas systems.

207

Page 225: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

8. IHS Harmony results with constant field SRV half-length of 87 m and matrix permeability of

6.5 x10-5 mD produce similar SRV permeability during production to the SRV permeability

from finite element analysis during hydraulic fracturing.

9. The effective permeability of the SRV during hydraulic fracturing from the FEA is estimated

to be of order of several tens of Darcies (23.4 D).

10. The SRV effective permeability during production from the new nonlinear PDE theory is of

the order of several tens of Darcies and it varies both spatially and temporally from 23.4 D (at

early production time and SRV boundary) to 0.0007 D (at late production time). This reveals

that the closure of the hydraulic fracture network occurs and the effect on permeability is

significant.

11. IHS Harmony analysis only produces one value of SRV effective permeability for all

production time and SRV distance. The SRV effective permeability from the best case of IHS

Harmony with constant field SRV half-length of 87 m and matrix permeability of 6.5 x10-5

mD is 13.5 D for PSS model and 22.3 D for sticks, slabs, and cubes models.

6.2 Recommendations

The recommendations for future research are:

1. Conduct core testing from a nearby well of Glauconitic Formation in Hoadley, AB to determine

the rock mechanical and the natural fractures properties for better and more accurate study.

2. Apply the finite element analysis on different reservoirs that have different values of Young’s

modulus and SRV dimensions to determine their impact on the SRV effective permeability.

3. Build our own elements in Abaqus to code cohesive elements to model hydraulic fractures to

implicitly model coupled pressure/deformation to model progressive damage of mechanical

strength, hydraulic conductivity, flow of fracturing fluid during opening fracture, and

permeability as a function of distance and time during hydraulic fracturing

4. Model hydraulic fracture propagation when intersecting a natural fracture as a function of

hydraulic fracture propagation angle, fracture aperture, injection pressure, proppant diameter,

and in-situ stresses during hydraulic fracturing. The model uses the Abaqus code mentioned in

recommendation number 3.

208

Page 226: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

5. Apply the new SRV dimensions model to other fields that have nearby old wellbores (depleted

zones) to observe if these have the same impact on the SRV growth and similar reservoir

geomechanic properties and hydraulic fracture injection parameters.

6. Optimize the new-analytical SRV dimensions model with integrating other reservoir

geomechanic properties and multi-phase hydraulic fracture injection fluid.

7. Model nonlinear PDE diffusivity using Matlab nonlinear PDE code editor on a 3D domain and

using non-constant boundary condition wellbore pressure to optimize the field gas flow rate

match.

8. Couple the nonlinear PDE diffusivity model with deformation and dual porosity models using

Matlab nonlinear PDE code editor to have an integrated models compare to IHS Harmony.

209

Page 227: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

Appendix A: Matlab nonlinear parabolic PDE diffusivity

codes % Problem Parameters using average viscosity and average Z. All units in % oilfield. % SRV and gas properties miu=0.0198319; % gas viscosity in cP, using Lee, Gonzales and Eakin equation http://checalc.com/solved/gasVisc.html at gas gravity 0.8, P=2900 psi (20MPa), T=158 F(70C), N2=0.05 mole %, CO2=0.0218 mole %, ,H2S=0 mole %. por=0.16; % porosity after hydraulic fracturing ct=4.024E-04; % ct in 1/psi that matches initial flow rates for por=0.16 S=7043.812016; % Carman-Kozeny specific surface, m2/m3 % Geometry and Mesh % For SRV rectangle, the geometry and mesh are defined as shown below: length= 570.866; % SRV length, ft (174 m) width= 196.85; % SRV width, ft (60 m) % For wellbore production rectangle, the geometry and mesh are defined as shown below: Wlength= 3.28084; % wellbore producing port length, ft (1 m) Wwidth= 3.28084; % wellbore producing port width, ft (1 m) % Define the rectangle by giving the 4 x-locations followed by the 4 % y-locations of the corners. % Rectangle is code 3, 4 sides, followed by x-coordinates and then y-coordinates % Rectangle SRV geometry R1 = [3 4 0 width width 0 0 0 length length]'; % Rectangle wellbore production edges geometry R2=[3 4 96.78458 100.06542 100.06542 96.78458 283.79458 283.79458 287.08 287.08]'; % Pad R2 with zeros to enable concatenation with R1 and R2 geom = [R1,R2]; % Names for the two geometric objects ns = (char('R1','R2'))'; % Set formula sf='R1-R2'; % Use this example: Deflection of a Piezoelectric Actuator % Create geometry g = decsg(geom,sf,ns); % View geometry % Plot the geometry and display the edge labels for use in the boundary % condition definition. figure;

210

Page 228: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

pdegplot(g, 'edgeLabels', 'on'); axis([0 600 0 600]); % in ft title 'Geometry With Edge Labels Displayed'; % Create the triangular mesh on the rectangular with approximately % ten elements in each direction. hmax = 16.4; % element sizehmax = 5 m (16.4 ft) within SRV, no need to use refinemesh [p, e, t] = initmesh(g, 'Hmax', hmax); figure; pdeplot(p,e,t); axis([0 600 0 600]); % in ft title 'SRV With Triangular Element Mesh' xlabel 'X-coordinate, feet' ylabel 'Y-coordinate, feet' % Boundary Conditions % Use PDE problem setup-BC-Examples-Applying constant BC % Create a pde entity for a PDE with a single dependent variable numberOfPDE = 1; pb = pde(numberOfPDE); % Scalar problem % Create a geometry entity pg = pdeGeometryFromEdges(g); % Create geometry object % BC: PDE problem setup-BC-examples-applying constant BC % Boundary conditions for P wellbore assumed to be equal average production % during 8 months after stabilize % welllbore P 500.23516 psi (3.449 MPa) need to convert P to u at wellbore bc3 = pdeBoundaryConditions(pg.Edges(3),'u',16224497.3); bc4 = pdeBoundaryConditions(pg.Edges(4),'u',16224497.3); bc5 = pdeBoundaryConditions(pg.Edges(5),'u',16224497.3); bc8 = pdeBoundaryConditions(pg.Edges(8),'u',16224497.3); % Put all the boundary conditions into a problem container. pb.BoundaryConditions =[bc3,bc4,bc5,bc8]; % All boundary conditions. % Solve the parabolic PDE with these boundary conditions with the PDE % nonlinear coefficients d and c, the nonlinear solver pdenonlin must be used to obtain the solution. % Definition of PDE Coefficients % The expressions for the coefficients required by the PDE toolbox can % easily be identified by comparing the equation with the scalar parabolic % equation in the PDE toolbox documentation % Permeability as a function of pressure % c=k, PDE nonlinear c coefficient, assumes kx=ky=kz, cx=cy=cz % c-coefficient using pseudo p c=sprintf('((%g.*(1-%g.*(2900.75-0.124190785.*(u.^0.5)))).^3)./(5.*(%g.^2).*(1-(%g.*(1-%g.*(2900.75-0.124190785.*(u.^0.5)))).^2)).*(10.^10)',por,ct,S,por,ct); % it works, c is in m2 originally needs to convert to mD by multiply 10^15. U0=2900.75 psi (20 MPa)

211

Page 229: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

%c equation above is following the example of "Nonlinear Heat Transfer In a %Thin Plate" when a-coefficient is a function of u a= 0; % if a is not 0 the solution plot profile is more realistic f= 0; % if f is not 0 the solution plot profile is more realistic % d-coefficient using pseudo p d= sprintf('(%g.*(1-%g.*(2900.75-0.124190785.*(u.^0.5))).*%g.*%g)./0.00005', por, ct, miu, ct); % we use 0.00005 to convert to field units, d=porf*miu*ct units are second. % Transient solution endTime = 6300; % Production in 828 days, hours tlist = 0:157.5:endTime; % Set the initial pressure of all nodes to initial reservoir pressure, 9 % MPa % Initial condition for pore pressure definition u0 =5.45562928746E+08; % initial pressure of all nodes average pressure along SRV after HF (start of production) in psi (20 MPa) rtol = 1.0e-3; atol = 1.0e-4; % The transient solver parabolic automatically handles both linear % and nonlinear problems, such as this one. u = parabolic(u0, tlist, pb,p,e,t,c,a,f,d); figure; % Post processing Pressure Pressure=0.124190785.*(u.^0.5); plot(tlist, Pressure(:,:)); grid on title 'Pressure as a Function of Time' xlabel 'Time, hours' ylabel 'Pressure, psi' figure; % Plot Pressure pdeplot(p, e, t, 'xydata', Pressure(:,end), 'contour', 'on', 'colormap', 'jet', 'mesh', 'on'); title(sprintf('Pressure In SRV (psi) (%d hours)\n', ... tlist(1,end))); % tlist(1,end), means end means final time or 19872 hours it is the same with tlist(end,end), if tlist(1,1) means time 1 or 0 hours it is the same with tlist(end,1) xlabel 'X-coordinate, feet' ylabel 'Y-coordinate, feet' axis ([0 600 0 600]); figure;

212

Page 230: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

% Porosity porf=sprintf('%g.*(1-%g.*(2900.75-0.124190785.*(u.^0.5)))', por, ct); % porosity final=por*(1-ct*(u0-u)) in fraction porf=por*(1-ct*(2900.75-0.124190785.*(u.^0.5))); % Plot porosity pdeplot(p, e, t, 'xydata', porf(:,end), 'contour', 'on', 'colormap', 'jet', 'mesh', 'on'); title(sprintf('Porosity In SRV (fraction) (%d hours)\n', ... tlist(1,end))); % tlist(1,end), means end means final time or 19872 hours it is the same with tlist(end,end), if tlist(1,1) means time 1 or 0 hours it is the same with tlist(end,1) xlabel 'X-coordinate, feet' ylabel 'Y-coordinate, feet' axis ([0 600 0 600]); figure; % Plot porosity vs time plot(tlist, porf(:,:)); title 'Porosity As a Function of time' xlabel 'Time, hours' ylabel 'Porosity, fraction' figure; % Plot permeability vs time k=((porf.^3)./(5.*S.*S.*((1-porf).^2))).*(10.^15); plot(tlist,k(:,:)); title 'Permeability As a Function of time' xlabel 'Time, hours' ylabel 'Permeability, mD' figure; % Plot Permeability vs distance at particular time pdeplot(p, e, t, 'xydata', k(:,end), 'contour', 'on', 'colormap', 'jet', 'mesh', 'on'); title(sprintf('Permeability In SRV (mD) (%d hours)\n', ... tlist(1,end))); % tlist(1,end), means end means final time or 19872 hours it is the same with tlist(end,end), if tlist(1,1) means time 1 or 0 hours it is the same with tlist(end,1) xlabel 'X-coordinate, feet' ylabel 'Y-coordinate, feet' axis ([0 600 0 600]); % Calculate dp/dx and dp/dy at each time separately [ux1,uy1] = pdegrad(p,t,u(:,1)); unx1 = pdeprtni(p,t,ux1); dpdx1=0.06209.*(u(:,1).^(-0.5)).*unx1; uny1 = pdeprtni(p,t,uy1); dpdy1=0.06209.*(u(:,1).^(-0.5)).*uny1; [ux2,uy2] = pdegrad(p,t,u(:,2)); unx2 = pdeprtni(p,t,ux2);

213

Page 231: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

dpdx2=0.06209.*(u(:,2).^(-0.5)).*unx2; uny2 = pdeprtni(p,t,uy2); dpdy2=0.06209.*(u(:,2).^(-0.5)).*uny2; [ux3,uy3] = pdegrad(p,t,u(:,3)); unx3 = pdeprtni(p,t,ux3); dpdx3=0.06209.*(u(:,3).^(-0.5)).*unx3; uny3 = pdeprtni(p,t,uy3); dpdy3=0.06209.*(u(:,3).^(-0.5)).*uny3; [ux4,uy4] = pdegrad(p,t,u(:,4)); unx4 = pdeprtni(p,t,ux4); dpdx4=0.06209.*(u(:,4).^(-0.5)).*unx4; uny4 = pdeprtni(p,t,uy4); dpdy4=0.06209.*(u(:,4).^(-0.5)).*uny4; [ux5,uy5] = pdegrad(p,t,u(:,5)); unx5 = pdeprtni(p,t,ux5); dpdx5=0.06209.*(u(:,5).^(-0.5)).*unx5; uny5 = pdeprtni(p,t,uy5); dpdy5=0.06209.*(u(:,5).^(-0.5)).*uny5; [ux6,uy6] = pdegrad(p,t,u(:,6)); unx6 = pdeprtni(p,t,ux6); dpdx6=0.06209.*(u(:,6).^(-0.5)).*unx6; uny6 = pdeprtni(p,t,uy6); dpdy6=0.06209.*(u(:,6).^(-0.5)).*uny6; [ux7,uy7] = pdegrad(p,t,u(:,7)); unx7 = pdeprtni(p,t,ux7); dpdx7=0.06209.*(u(:,7).^(-0.5)).*unx7; uny7 = pdeprtni(p,t,uy7); dpdy7=0.06209.*(u(:,7).^(-0.5)).*uny7; [ux8,uy8] = pdegrad(p,t,u(:,8)); unx8 = pdeprtni(p,t,ux8); dpdx8=0.06209.*(u(:,8).^(-0.5)).*unx8; uny8 = pdeprtni(p,t,uy8); dpdy8=0.06209.*(u(:,8).^(-0.5)).*uny8; [ux9,uy9] = pdegrad(p,t,u(:,9)); unx9 = pdeprtni(p,t,ux9); dpdx9=0.06209.*(u(:,9).^(-0.5)).*unx9; uny9 = pdeprtni(p,t,uy9); dpdy9=0.06209.*(u(:,9).^(-0.5)).*uny9; [ux10,uy10] = pdegrad(p,t,u(:,10)); unx10 = pdeprtni(p,t,ux10); dpdx10=0.06209.*(u(:,10).^(-0.5)).*unx10; uny10 = pdeprtni(p,t,uy10); dpdy10=0.06209.*(u(:,10).^(-0.5)).*uny10; [ux11,uy11] = pdegrad(p,t,u(:,11)); unx11 = pdeprtni(p,t,ux11); dpdx11=0.06209.*(u(:,11).^(-0.5)).*unx11; uny11 = pdeprtni(p,t,uy11); dpdy11=0.06209.*(u(:,11).^(-0.5)).*uny11; [ux12,uy12] = pdegrad(p,t,u(:,12)); unx12 = pdeprtni(p,t,ux12); dpdx12=0.06209.*(u(:,12).^(-0.5)).*unx12; uny12 = pdeprtni(p,t,uy12); dpdy12=0.06209.*(u(:,12).^(-0.5)).*uny12; [ux13,uy13] = pdegrad(p,t,u(:,13)); unx13 = pdeprtni(p,t,ux13); dpdx13=0.06209.*(u(:,13).^(-0.5)).*unx13; uny13 = pdeprtni(p,t,uy13);

214

Page 232: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

dpdy13=0.06209.*(u(:,13).^(-0.5)).*uny13; [ux14,uy14] = pdegrad(p,t,u(:,14)); unx14 = pdeprtni(p,t,ux14); dpdx14=0.06209.*(u(:,14).^(-0.5)).*unx14; uny14 = pdeprtni(p,t,uy14); dpdy14=0.06209.*(u(:,14).^(-0.5)).*uny14; [ux15,uy15] = pdegrad(p,t,u(:,15)); unx15 = pdeprtni(p,t,ux15); dpdx15=0.06209.*(u(:,15).^(-0.5)).*unx15; uny15 = pdeprtni(p,t,uy15); dpdy15=0.06209.*(u(:,15).^(-0.5)).*uny15; [ux16,uy16] = pdegrad(p,t,u(:,16)); unx16 = pdeprtni(p,t,ux16); dpdx16=0.06209.*(u(:,16).^(-0.5)).*unx16; uny16 = pdeprtni(p,t,uy16); dpdy16=0.06209.*(u(:,16).^(-0.5)).*uny16; [ux17,uy17] = pdegrad(p,t,u(:,17)); unx17 = pdeprtni(p,t,ux17); dpdx17=0.06209.*(u(:,17).^(-0.5)).*unx17; uny17 = pdeprtni(p,t,uy17); dpdy17=0.06209.*(u(:,17).^(-0.5)).*uny17; [ux18,uy18] = pdegrad(p,t,u(:,18)); unx18 = pdeprtni(p,t,ux18); dpdx18=0.06209.*(u(:,18).^(-0.5)).*unx18; uny18 = pdeprtni(p,t,uy18); dpdy18=0.06209.*(u(:,18).^(-0.5)).*uny18; [ux19,uy19] = pdegrad(p,t,u(:,19)); unx19 = pdeprtni(p,t,ux19); dpdx19=0.06209.*(u(:,19).^(-0.5)).*unx19; uny19 = pdeprtni(p,t,uy19); dpdy19=0.06209.*(u(:,19).^(-0.5)).*uny19; [ux20,uy20] = pdegrad(p,t,u(:,20)); unx20 = pdeprtni(p,t,ux20); dpdx20=0.06209.*(u(:,20).^(-0.5)).*unx20; uny20 = pdeprtni(p,t,uy20); dpdy20=0.06209.*(u(:,20).^(-0.5)).*uny20; [ux21,uy21] = pdegrad(p,t,u(:,21)); unx21 = pdeprtni(p,t,ux21); dpdx21=0.06209.*(u(:,21).^(-0.5)).*unx21; uny21 = pdeprtni(p,t,uy21); dpdy21=0.06209.*(u(:,21).^(-0.5)).*uny21; [ux22,uy22] = pdegrad(p,t,u(:,22)); unx22 = pdeprtni(p,t,ux22); dpdx22=0.06209.*(u(:,22).^(-0.5)).*unx22; uny22 = pdeprtni(p,t,uy22); dpdy22=0.06209.*(u(:,22).^(-0.5)).*uny22; [ux23,uy23] = pdegrad(p,t,u(:,23)); unx23 = pdeprtni(p,t,ux23); dpdx23=0.06209.*(u(:,23).^(-0.5)).*unx23; uny23 = pdeprtni(p,t,uy23); dpdy23=0.06209.*(u(:,23).^(-0.5)).*uny23; [ux24,uy24] = pdegrad(p,t,u(:,24)); unx24 = pdeprtni(p,t,ux24); dpdx24=0.06209.*(u(:,24).^(-0.5)).*unx24; uny24 = pdeprtni(p,t,uy24); dpdy24=0.06209.*(u(:,24).^(-0.5)).*uny24; [ux25,uy25] = pdegrad(p,t,u(:,25));

215

Page 233: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

unx25 = pdeprtni(p,t,ux25); dpdx25=0.06209.*(u(:,25).^(-0.5)).*unx25; uny25 = pdeprtni(p,t,uy25); dpdy25=0.06209.*(u(:,25).^(-0.5)).*uny25; [ux26,uy26] = pdegrad(p,t,u(:,26)); unx26 = pdeprtni(p,t,ux26); dpdx26=0.06209.*(u(:,26).^(-0.5)).*unx26; uny26 = pdeprtni(p,t,uy26); dpdy26=0.06209.*(u(:,26).^(-0.5)).*uny26; [ux27,uy27] = pdegrad(p,t,u(:,27)); unx27 = pdeprtni(p,t,ux27); dpdx27=0.06209.*(u(:,27).^(-0.5)).*unx27; uny27 = pdeprtni(p,t,uy27); dpdy27=0.06209.*(u(:,27).^(-0.5)).*uny27; [ux28,uy28] = pdegrad(p,t,u(:,28)); unx28 = pdeprtni(p,t,ux28); dpdx28=0.06209.*(u(:,28).^(-0.5)).*unx28; uny28 = pdeprtni(p,t,uy28); dpdy28=0.06209.*(u(:,28).^(-0.5)).*uny28; [ux29,uy29] = pdegrad(p,t,u(:,29)); unx29 = pdeprtni(p,t,ux29); dpdx29=0.06209.*(u(:,29).^(-0.5)).*unx29; uny29 = pdeprtni(p,t,uy29); dpdy29=0.06209.*(u(:,29).^(-0.5)).*uny29; [ux30,uy30] = pdegrad(p,t,u(:,30)); unx30 = pdeprtni(p,t,ux30); dpdx30=0.06209.*(u(:,30).^(-0.5)).*unx30; uny30 = pdeprtni(p,t,uy30); dpdy30=0.06209.*(u(:,30).^(-0.5)).*uny30; [ux31,uy31] = pdegrad(p,t,u(:,31)); unx31 = pdeprtni(p,t,ux31); dpdx31=0.06209.*(u(:,31).^(-0.5)).*unx31; uny31 = pdeprtni(p,t,uy31); dpdy31=0.06209.*(u(:,31).^(-0.5)).*uny31; [ux32,uy32] = pdegrad(p,t,u(:,32)); unx32 = pdeprtni(p,t,ux32); dpdx32=0.06209.*(u(:,32).^(-0.5)).*unx32; uny32 = pdeprtni(p,t,uy32); dpdy32=0.06209.*(u(:,32).^(-0.5)).*uny32; [ux33,uy33] = pdegrad(p,t,u(:,33)); unx33 = pdeprtni(p,t,ux33); dpdx33=0.06209.*(u(:,33).^(-0.5)).*unx33; uny33 = pdeprtni(p,t,uy33); dpdy33=0.06209.*(u(:,33).^(-0.5)).*uny33; [ux34,uy34] = pdegrad(p,t,u(:,34)); unx34 = pdeprtni(p,t,ux34); dpdx34=0.06209.*(u(:,34).^(-0.5)).*unx34; uny34 = pdeprtni(p,t,uy34); dpdy34=0.06209.*(u(:,34).^(-0.5)).*uny34; [ux35,uy35] = pdegrad(p,t,u(:,35)); unx35 = pdeprtni(p,t,ux35); dpdx35=0.06209.*(u(:,35).^(-0.5)).*unx35; uny35 = pdeprtni(p,t,uy35); dpdy35=0.06209.*(u(:,35).^(-0.5)).*uny35; [ux36,uy36] = pdegrad(p,t,u(:,36)); unx36 = pdeprtni(p,t,ux36); dpdx36=0.06209.*(u(:,36).^(-0.5)).*unx36;

216

Page 234: On Hydraulic Fracturing of Tight Gas Reservoir Rock...On Hydraulic Fracturing of Tight Gas Reservoir Rock by Belladonna Maulianda A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

uny36 = pdeprtni(p,t,uy36); dpdy36=0.06209.*(u(:,36).^(-0.5)).*uny36; [ux37,uy37] = pdegrad(p,t,u(:,37)); unx37 = pdeprtni(p,t,ux37); dpdx37=0.06209.*(u(:,37).^(-0.5)).*unx37; uny37 = pdeprtni(p,t,uy37); dpdy37=0.06209.*(u(:,37).^(-0.5)).*uny37; [ux38,uy38] = pdegrad(p,t,u(:,38)); unx38 = pdeprtni(p,t,ux38); dpdx38=0.06209.*(u(:,38).^(-0.5)).*unx38; uny38 = pdeprtni(p,t,uy38); dpdy38=0.06209.*(u(:,38).^(-0.5)).*uny38; [ux39,uy39] = pdegrad(p,t,u(:,39)); unx39 = pdeprtni(p,t,ux39); dpdx39=0.06209.*(u(:,39).^(-0.5)).*unx39; uny39 = pdeprtni(p,t,uy39); dpdy39=0.06209.*(u(:,39).^(-0.5)).*uny39; [ux40,uy40] = pdegrad(p,t,u(:,40)); unx40 = pdeprtni(p,t,ux40); dpdx40=0.06209.*(u(:,40).^(-0.5)).*unx40; uny40 = pdeprtni(p,t,uy40); dpdy40=0.06209.*(u(:,40).^(-0.5)).*unx40; [ux41,uy41] = pdegrad(p,t,u(:,41)); unx41 = pdeprtni(p,t,ux41); dpdx41=0.06209.*(u(:,41).^(-0.5)).*unx41; uny41 = pdeprtni(p,t,uy41); dpdy41=0.06209.*(u(:,41).^(-0.5)).*uny41;

217


Recommended